You are on page 1of 571

School of Civil and Environmental Engineering

Faculty of Engineering and Information Technology


University of Technology Sydney
Sydney, Australia

COMPARATIVE STUDY OF THE LONG-TERM DEFLECTION OF


CONVENTIONAL AND SELF-COMPACTING CONCRETE WITH
LIGHT-WEIGHT CONCRETE SLABS

By
Behnam Vakhshouri
BSc (Civil Eng.), MSc (Structural Eng.)

A thesis submitted in fulfillment of


the requirements for the degree of
Doctor of Philosophy

Principal supervisor: Dr. Shami Nejadi


Co-Supervisor: Dr. Emre Erkmen

June 2017
CERTIFICATE OF AUTHORSHIP/ORIGINALITY

I certify that the work in this thesis has not previously been submitted for a degree nor has it

been submitted as part of requirements for a degree except as fully acknowledged within the

text. I also certify that the thesis has been written by me. Any help that I have received in my

research work and the preparation of the thesis itself has been acknowledged. In addition, I

certify that all information sources and literature used are indicated in the thesis.

Signature

Behnam Vakhshouri
June 2017
AKNOWLEDGEMENTS

I want to thank everyone who assisted in this great endeavor; for without you, I would
not have been able to accomplish so much in so little time. I would like to express my
special appreciation and thanks to my supervisor Dr. SHAMI NEJADI, for his priceless
support, guidance, and direction in my educational and career development. I would like
to thank DR. EMRE ERKMEN for his great supports and directions especially in the
numerical part of the study. I would also like to thank the staff in the concrete
laboratory of the University of Technology Sydney for their continuous help to conduct
my experimental part of the research.

I would like to express my appreciation to the ones who have been and will always be
by my side. I would also like to thank all of my friends who supported me in writing,
and incented me to strive towards my goal.

I would like to thank my family for their continuous encouragement. A special thanks to
my father-in-law, DR. MOHAMMADREZA SHADAMANAMAN for all kinds of his
support. I am so very grateful for my daughter MELORIN who was always waiting at
the door to give dad hugs and to brighten my day. And finally, I would like to thank my
beloved wife, ZAHRA, who spent sleepless nights with and was always my support in
every moment.

I dedicate this thesis to the spirits of my mother and father who sacrificed their life
to brighten my life.

Behnam Vakhshouri
June 2017
Notations
The symbols used in this study. Including their definitions, are listed below.

AASHTO American Association of State Highway and Transportation Officials


ACI American Concrete Institute
AEA Air-Entraining Agent
ALWAC All Light-Weight Aggregate Concrete
AS Australian standard
ASTM American Society for Testing and Materials
Acr Area of section at first cracking
Ae Effective section area under direct tension
Ag Gross cross section area
Ast Area of tensile steel bar

Commercial name for the spherical-shaped polystyrene beads with hydrophilic


BST
type chemical coating

CC Conventional concrete
CS Compressive Strength
CS.CUBE Compressive strength of cube specimen
CSH Calcium Silicate Hydrates
CS.CYL Compressive strength of cylindrical specimen
CSSC Compressive stress-strain curve
Cyl. Cylinder
Cb Concrete cover of bar to the soffit of member
Cm Smaller of clear bottom cover in reinforced concrete
c/p Cement to powder ratio
Cs Concrete cover of bar to the side surface of member

DL Dead load
D-SCC Self-compacting concrete containing steel fibre
DS-SCC Self-compacting concrete containing hybrid fibre (steel and Polypropylene fibre)
‫ܦ‬଴௘௟ Initial (undamaged) elastic stiffness of the material
d Effective depth of tensile bar
dia. Diameter
db Bar diameter

ECC Engineered Cementitious Composite


EI Bending stiffness
EPS Expanded Poly-Styrene
EPS-LWC Light Weight Concrete containing Expanded Poly-Styrene Beads
Ec Modulus of elasticity of concrete
a
Ec-14 Modulus of elasticity of concrete at age 14 days
Ec-28 Modulus of elasticity of concrete at age 28 days
Ef Modulus of elasticity of fibre
Ep Final slope of the compressive stress-strain curve
Es Modulus of elasticity of steel bars
Initial slope of the compressive stress-strain curve

F Axial tension in pull out test


FEA Finite Element Analysis
FHWA Federal Highway Administration
FRP Fibre reinforced polymer
FRSCC Fibre reinforced self-compacting concrete
fck Specified compressive strength of concrete
fcm Mean concrete cylinder compressive strength
fct Splitting tensile strength of concrete
fct-28 Splitting tensile strength of concrete at age 28 days
fcu Specified characteristic cube compressive strength
Fd.ef Effective design load, per unit area
fr Modulus of rupture (flexural tensile strength)
frel Relative compressive strength
fr-28 Modulus of rupture at age 28 days
fs estimated design service stress in tension reinforcement
ft Splitting tensile strength
ft. Foot
fy yield strength of reinforcement
fyt Yield stress of tensile bar
f (t) Time-function parameter
݂௖ᇱ Compressive strength of concrete
ᇱ௘௙
݂௧ Effective tensile strength derived from the biaxial failure function of concrete

GGBFS Ground Granulated Blast Furnace Slag


gr Gram
GPa Giga Pascal
Gc Absorbed energy per unit volume of the cylindrical specimens under compression
Absorbed energy per unit volume of the cylindrical specimens under compression
Gc-28
at age 28 days
Gf Fracture energy needed to create a unit area of stress-free crack
Average energy required for unit crack propagation from crack initiation until any
GFi
time

HPFRC High-Performance Fibre Reinforced Concrete


HRWRA High Range Water Reducing Admixture
HSLWC High-Strength Light-Weight Concrete
HSNWC High-Strength Normal-Weight Concrete
hr Rib height in steel bar
b
Icr Moment of inertia of section at first cracking
Ie Effective moment of inertia
Ig gross moment of inertia of uncracked reinforced concrete section
Im Average moment of inertia of section in span length

J (t, t0) Compliance function of concrete

LEFM Linear Elastic Fracture Mechanics

l Span length
lb pound
lb/ft3 Pound per cubic feet
3
Lit/m Litre per cubic metre
LL1 Loading level 1 (Lower loading level)
LL2 Loading level 2 (Upper loading level)
LVDT Linear variable differential transformer
LWC Light weight concrete
LWCC Light-Weight Ceramic Concrete
Lcr Cracked length
ld Embedment length of steel bar in concrete
Lef Effective span length
ln Length of clear span length measured face to face of supports

kips Imperial unit of force equal to 1000 pounds-force


Kg Kilo grams
kg/m3 Kilo gram per cubic meters
KN Kilo Newton
KPa Kilo Pascal
Ksi Size effect factor
kcs Coefficient of creep and shrinkage in long-term deflection calculations
Kp Final slope of the compressive stress-strain curve

max Maximum
min Minimum
ml/kg Millilitre per kilo gram
mm Millimetre
mm/mm 0K Millimetre / millimetre grade Kelvin
MM Moisture migration
MoE Modulus of elasticity
MoR Modulus of rupture (flexural tensile strength)
MPa Mega Pascal
Ma bending moment due to servivce loads (service moment)
c
Ratio of service moment to the ultimate moment capacity of reinforced concrete
Ma / M u
section
Mcr Cracking moment of reinforced concrete section
Ratio of cracking moment to the ultimate moment capacity of reinforced concrete
Mcr / Mu
section
Mmax Maximum bending moment at critical section
Mu Ultimate bending moment of reinforced concrete section
m1 Weak axis of the material
m2 Strong axis of the material
m3 cubic meters
M–κ Moment- curvature

N Vector of applied loads


N-SCC Normal self-compacting concrete
N/mm2 Newton per square millimetre
N(t) Resisting force (reaction) in restrained shrinkage test

OPC Ordinary Portland Cement


OPKS Oil Palm Kernel Shell

pcf Pound per cubic inch


PET Poly-Ethylene Terephthalate
PFA Pulverised Fuel Ash
PL60-11-3L Type of strain gauge to record the strain variation in concrete surface
psi Pounds per square inch
Pcr Tension force at first cracking under direct tension

RC Reinforced concrete
RH Relative Humidity
Reunion Internationale des Laboratoires et Experts des Materiaux (International
RILEM Union of Laboratories and Experts in Construction Materials, Systems, and
Structures)
RMS Road and maritime services
RTA Regional Transportation Authority
RVDT Rotary variable differential transformer
rec Reduction factor of the compressive strength
Relec Electrical resistivity
ret Reduction factor of the tensile strength
Rhm Rebound value (Schmidt hammer)
Rr Ratio of the rib height to the rib spacing in ribbed bar

S Bar spacing
s Slip of steel bar in concrete
SCC Self-compacting concrete
SFRC Steel Fibre Reinforced Concrete
SHCC Strain Hardening Cementitious Composites
d
SLWAC Sand Light-Weight Aggregate Concrete
SP Super Plasticiser
S-SCC Self-compacting concrete containing Polypropylene fibre
STS Splitting tensile strength
S0 bond transfer length in short-term loading
ܵ଴‫כ‬ bond transfer length in long-term loading
sm Slip of steel bar in concrete at peak bond stress
Sr Final average crack spacing
sr Rib spacing in steel bar

t Time
Tmax Ultimate bond force
tc Age of concrete where moist-curing ends in days

UDL Uniformly distributed load


UHPFRC Ultra High-Performance Fibre Reinforced Concrete
UNSW University of New South Wales
UPV Ultrasonic Pulse Velocity
UTS Universtity of Technology Sydney
Ui Elastic strain energy

Ratio of Volume to drying surface area of concrete specimen (Hypothetical


V/S
thickness)
vol.% Volumetric percentage
Vair Entrained air content by volume
Vp Loading rate in pull out test
VΔW Rate of weight loss of shrinkage specimen with time

w Crack width
w/c Water/cementitious material ratio
WFLA6-11 Type of strain gauge to record the strain variation in steel bars
wa Applied service load
wc Crack width at the complete release of stress
Wi Applied work of external force P
wmax-t Maximum crack width at time t
W0 Initial weight of shrinkage specimen

γ Density
γRH Coefficient of relative humidity in creep coefficient calculation
γVS Coefficient of hypothetical thickness in creep coefficient calculation
γt0 Coefficient of loading time in creep coefficient calculation
γxy Shear strain component in plane x and direction y
γyz Shear strain component in plane y and direction z
γzx Shear strain component in plane x and direction z

e
ΔU Vector of current nodal displacement
Δcr Deflection due to creep effect
Δe Flexural elastic deflection of slab
Δe-14 Flexural elastic deflection of slab based on the modulus of elasticity at 14 days
Δe-28 Flexural elastic deflection of slab based on the modulus of elasticity at 28 days
δi Crack opening displacement
Δinc Incremental flexural deflection
Δins Instantaneous flexural deflection
Δins-14 Flexural instantaneous deflection of slab subjected to loading at age 14 days
Δi-t Flexural deflection of point i at time t
Ratio of flexural instantaneous deflection to the elastic deflection based on the
(Δins / Δe)14
modulus of elasticity at 14 days
Ratio of flexural instantaneous deflection at 14 days to the elastic deflection of
Δins-14 / Δe-28
slab based on the modulus of elasticity at 28 days
Δlong Long-term flexural deflection
Δlong / Δins Ratio of long-term to instantaneous flexural deflection
Δsh Deflection due to shrinkage (autogeneous and chemical shrinkage)
Δsus Time-dependent deflection under sustained loading
Δtmp Temperature induced deflection
Δtot Total flexural deflection
δκ0.ts Average instantaneous curvature
ΔW Weight loss of shrinkage specimen
ΔW/W0 Ratio of lost weight to the initial weight of shrinkage specimen
∆⁄Lef Flexural deflection limit of beam and slab
Δ1 Vertical deflection of slab at mid-span
Δ2 vertical deflection of slab at the border of high moment region
Δe1 Calculated elastic deflection of slab at mid-span
Δe2 Calculated elastic deflection of slab at the border of high moment region
Δ1-SS Flexural deflection at midspan of simply supported slab
Δ1-FE Flexural deflection at midspan of fixed supports slab
Δ2-FE Flexural deflection at the border of high moment region in fixed supports slab
Δ2-SS Flexural deflection at the border of high moment region in simply supported slab

ε Strain
ߝ௖ᇱ Corresponding strain to the maximum compressive strength of concrete
εcr Creep strain
εcsd.b Basic drying shrinkage strain
‫כ‬
ߝ௖௦ௗǤ௕ Final drying basic shrinkage strain
‫כ‬
ߝ௖௦௘ Final autogeneous shrinkage strain
εe Elastic strain
εp Plastic strain
εs Tensile strain in steel bar
εsh Shrinkage strain
εz Normal strain component in z direction
εcse Autogeneous shrinkage strain
εcsd Drying shrinkage strain
f
εs2 Average strain in a tensioned member
εy Ultimate strain of tensile steel bar
εu Ultimate strain of the compressive stress-strain curve
εo Corresponding strain to the maximum compressive strength of concrete

ξ Scalar stiffness degradation variable

κ Curvature
κsh Shrinkage-induced curvature
(κsh)cr Shrinkage-induced curvature on a previously cracked cross section
(κsh)uncr shrinkage-induced curvature in uncracked cross-section

λ Crack spacing

με Micro strain

ν Poison ratio

ρ Ratio of tensile steel bar in the section


ρb Ratio of tensile steel bar in the section at balance condition

σ0 Reference compressive stress of concrete


σe Effective stress
σ-ε Stress-strain
σc Compressive Strength of concrete
σz Normal stress component in direction z
σy Normal stress component in direction y
σx Normal stress component in direction x
σc1 Concrete stress away from the crack

σs1 Steel stress away from the crack


σs2 Steel stress at the crack
‫כ‬
ߪ௦ଵ Steel stress away from the crack after all shrinkage cracking
‫כ‬
ߪ௦ଶ Steel stress at the crack after all shrinkage cracking

τ bond stress
τb-average Average bond stress to describe the force transfer between steel bar and concrete
τmax Peak bond stress
τxz Shear stress component in plane x normal to z direction
τyz Shear stress component in plane y normal to z direction
τu Ultimate bond stress

ϕcc Creep coefficient


φ (t,t0) Creep coefficient
g
ϕc.b Basic creep coefficient
ϕu Ultimate creep coefficient

0
K Degree Kelvin

2D Two- dimensional
3D Three-dimensional

% Percentage

h
ABSTRACT

Long-term deflection of concrete slabs is often the main governing design criteria to determine
the appropriate thickness of the slabs to meet the required serviceability limit state. Due to the
nonlinear nature of the material, predicting the long-term deflection of concrete structures is
complex. Also, the complicated behaviour of concrete in the presence of reinforcement makes
the problem more challenging to study.

In general, codes of practice give an overall estimation of the long-term deflection as a


multiplier of the short-term or elastic deflection of concrete structures. Such estimation
sometimes may lead to an entirely wrong prediction of deflection and consequently unsafe,
unrealistic and unprofitable concrete design and construction.

Recent developments in concrete technology have led to produce new construction materials by
significant strength and performance features. Many of these developments are engineered
solutions to technical and commercial problems by either improvement of the current practices
or overcoming of limitations in the existing construction technology. Lightweight and self-
compacting concrete are the main two innovative concrete types used in the construction
industry along with conventional concrete. This study examines and evaluates the time-
dependent deflection of reinforced concrete one-way slabs made of two concrete types versus
the conventional concrete.

The presented investigation comprises both experimental and analytical components. The
experimental part consists of laboratory investigation of the time-dependent flexural deflection
and monitoring of the strains in reinforced concrete slabs. Since lightweight concrete is very
prevalent in building construction in Australia, and there are limited studies on its long-term
behaviour, the current study examines the lightweight concrete slabs identical to the previously
tested conventional concrete slabs (Nejadi, 2005) and self-compacting concrete slabs (Aslani,
2014) subjected to the identical long-term loading.

In addition, an analytical study consists of verification of the recorded experimental data and
parametric study of the effective factors in the time-dependent behaviour and deflection of the
conventional and self-compacting concrete slabs are conducted. Recently developed high-
capable ATENA software is utilized in the numerical analysis. Load-deflection behaviour of the
slabs under short-term loading is also recorded and compared with those of self-compacting and
conventional concrete slabs.

1
Finally, the results of this study are used to evaluate and verify the existing and proposed
models associated with the parameters that affect the flexural deflection of reinforced concrete
slabs with different types of concrete.

2
Table of contents

Chapter 1: Introduction

1.1. BACKGROUND ............................................................................................................................ 1


1.2. STATE OF THE PROBLEM.......................................................................................................... 2
1.3. OBJECTIVES AND SCOPE OF THE RESEARCH ...................................................................... 4
1.4. METHODOLOGY .......................................................................................................................... 5
1.5. LAYOUT OF THE THESIS ........................................................................................................... 6
1.6. SUMMARY .................................................................................................................................... 7
REFERENCES ............................................................................................................................................. 8

Chapter 2: Literature review

2.1. BACKGROUND AND DEVELOPMENT OF LIGHTWEIGHT CONCRETE ............................ 9


2.2. MIX DESIGN OF STRUCTURAL LIGHTWEIGHT CONCRETE ............................................ 14
2.3. DEFINITION OF LIGHTWEIGHT CONCRETE ....................................................................... 16
2.4. CLASSIFICATION OF LIGHTWEIGHT CONCRETE .............................................................. 19
2.5. ADVANTAGES AND DISADVANTAGES OF LIGHTWEIGHT CONCRETE ....................... 20
2.6. FRESH PROPERTIES OF LIGHTWEIGHT CONCRETE ......................................................... 24
2.7. MECHANICAL PROPERTIES OF LIGHTWEIGHT CONCRETE ........................................... 28
2.8. FRACTURE ENERGY ................................................................................................................. 35
2.9. SHRINKAGE AND CREEP OF LIGHTWEIGHT CONCRETE ................................................ 37
2.10. REAL-SCALE TIME-DEPENDENT DEFLECTION STUDIES ................................................ 42
2.11. SUMMARY .................................................................................................................................. 46
REFERENCES ........................................................................................................................................... 48

Chapter 3: Hardened concrete properties

3.1. INTRODUCTION ........................................................................................................................ 52


3.2. MANUFACTURING PROCESS OF EXPANDED POLYSTYRENE BEADS .......................... 54
3.3. EXPERIMENTAL DATA ............................................................................................................ 56
3.4. MECHANICAL PROPERTIES OF LIGHTWEIGHT CONCRETE CONTAINING
EXPANDED POLYSTYRENE BEADS ................................................................................................... 69
3.5. MECHANICAL PROPERTIES OF CONVENTIONAL AND SELF-COMPACTING
CONCRETE............................................................................................................................................... 90
3.6. COMPRESSIVE STRESS-STRAIN CURVE IN LIGHTWEIGHT, CONVENTIONAL AND
SELF-COMPACTING CONCRETE ......................................................................................................... 98
3.7. STEEL-CONCRETE BOND CHARACHTERISTICS .............................................................. 107
a
3.8. SUMMARY ................................................................................................................................ 110
REFERENCES ......................................................................................................................................... 112

Chapter 4: Time-dependent properties of concrete

4.1. INTRODUCTION ...................................................................................................................... 120


4.2. CREEP STRAIN ......................................................................................................................... 123
4.3. CREEP COEFFICIENT .............................................................................................................. 125
4.4. SHRINKAGE STRAIN .............................................................................................................. 126
4.5. EXISTING CREEP AND SHRINKAGE MODELS IN LIGHTWEIGHT, SELF-COMPACTING
AND CONVENTIONAL CONCRETE ................................................................................................... 127
4.6. SOME EXPERIMENTS ON CREEP AND SHRINKAGE OF LIGHTWEIGHT CONCRETE IN
THE LITERATURE ................................................................................................................................ 136
4.7. BOND TRANSFER LENGTH IN LIGHTWEIGHT CONCRETE ........................................... 136
4.8. EXISTING MODELS OF BOND-SLIP FOR LIGHTWEIGHT, CONVENTIONAL AND SELF-
COMPACTING CONCRETE ................................................................................................................. 149
4.9. SUMMARY ................................................................................................................................ 154
REFERENCES ......................................................................................................................................... 156

Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

5.1. INTRODUCTION ...................................................................................................................... 160


5.2. DEFLECTION CONTROL IN SERVICEABILITY DESIGN .................................................. 163
5.3. RATIO OF LONG-TERM TO SHORT-TERM DEFLECTION ................................................ 164
5.4. REAL-SCALE TIME-DEPENDENT DEFLECTION MONITORING IN LITERATURE ...... 165
5.5. EXPERIMENTAL MONITORING OF DEFLECTIONS IN LABORATORY CONDITIONS
FORM LITERATURE ............................................................................................................................. 167
5.6. MINIMUM THICKNESS PROVISIONS .................................................................................. 173
5.7. EFFECTIVE PARAMETERS IN INSTANTANEOUS DEFLECTION.................................... 175
5.8. EFFECTIVE PARAMETERS IN LONG-TERM DEFLECTION ............................................. 176
5.9. TENSION STIFFENING ............................................................................................................ 177
5.10. EFFECTIVE MOMENT OF INERTIA ...................................................................................... 187
5.11. EXISTING MODELS OF EFFECTIVE MOMENT OF INERTIA ........................................... 193
5.12. SUMMARY ................................................................................................................................ 195
CONFERENCES ..................................................................................................................................... 197

b
Chapter 6: Experimental program, Phase I- Properties of light-weight concrete containing
expanded polystyrene beads

6.1. INTRODUCTION ...................................................................................................................... 202


6.2. LIGHTWEIGHT CONCRETE CONTAINING EXPANDED POLYSTRYRENE BEADS ..... 204
6.3. MIXTURE PROPORTIONS ...................................................................................................... 210
6.4. PREPARATION AND CURING CONDITION OF THE TEST SPECIMENS ........................ 211
6.5. TEST METHODS OF THE HARDENED LIGHTWEIGHT CONCRETE ............................... 211
6.6. PROPERTIES OF THE FRESH LIGHTWEIGHT CONCRETE .............................................. 212
6.7. EXPERIMENTAL TEST RESULTS OF FRESH AND HARDENED LIGHTWEIGHT
CONCRETE............................................................................................................................................. 213
6.8. COMPARING THE MECHANICAL PROPERTIES OF LIGHTWEIGHT, SELF-
COMPACTING AND CONVENTIONAL CONCRETE ....................................................................... 259
6.9. SUMMARY ................................................................................................................................ 267
REFERENCES ......................................................................................................................................... 280

Chapter 7: Experimental program, Phase II, flexural short-term load-deflection

7.1. SHORT-TERM EXPERIMENTAL PROGRAM ....................................................................... 282


7.2. TEST PARAMETERS AND REINFORCEMENT LAYOUT................................................... 283
7.3. CONSTRUCTION OF THE SPECIMENS AND TEST PROCEDURE IN SHORT-TERM
LOAD-DEFLECTION TEST .................................................................................................................. 284
7.4. LOADING BLOCKS .................................................................................................................. 285
7.5. INSTRUMENTATION OF THE SLABSAND TEST SETUP .................................................. 286
7.6. LOAD-DEFLECTION BEHAVIOUR ....................................................................................... 288
7.7. PREVIOUS EXPERIMENTS ON LOAD-DEFLECTION BEHAVIOUR OF SLABS ............ 297
7.8. COMPARISON OF LOAD-DEFLECTION CURVES IN LIGHTWEIGHT, CONVENTIOANL
AND SELF-COMAPCTING CONCRETE SLABS ................................................................................ 302
7.9. SUMMARY ................................................................................................................................ 304
CONFERENCES ..................................................................................................................................... 308

Chapter 8: Experimental program, Phase II, flexural instantaneous deflection

8.1. INTRODUCTION ...................................................................................................................... 309


8.2. EXPERIMENTAL PROGRAM ................................................................................................. 311
8.3. EXPERIMENTAL INSTANTANEOUS DEFLECTION .......................................................... 315
8.4. DIFFERENCE BETWEEN THE ELASTIC AND INSTANTANEOUS DEFLECTION ......... 319
8.5. EFFECT OF LOADING LEVEL ............................................................................................... 321
8.6. PREVIOUS EXPERIMENTS ON THE INSTANTANEOUS DEFLKECTION OF SLABS ... 323

c
8.7. COMPARISON OF LOAD-DEFLECTION DIAGRAMS IN LIGHTWEIGHT, SELF-
COMPACTING AND CONVENTIONAL CONCRETE SLABS .......................................................... 325
8.8. PROPOSED ANALYTICAL MODEL ...................................................................................... 330
8.9. PROPOSED MODEL OF Ief IN SELF-COMPACTING CONCRETE SLABS ....................... 334
8.10. ROPOSED MODEL OF Ief IN LIGHTWEIGHT CONCRETE SLABS ................................... 335
8.11. SUMMARY ................................................................................................................................ 336
REFERENCES ......................................................................................................................................... 340

Chapter 9: Experimental program, Phase II, flexural long-term deflection

9.1. INTRODUCTION ...................................................................................................................... 343


9.2. EXPERIMENTAL PROGRAM ................................................................................................. 346
9.3. RATIO OF LONG-TERM TO INSTANTANEOUS DEFLECTION ........................................ 354
9.4. RATE OF TIME DEPENDENT DEFLECTION IN LIGHTWEIGHT CONCRETE SLABS... 356
9.5. CRACKING HISTORY OF THE LIGHTWEIGHT CONCRETE SLABS AT EARLY AGES OF
LOADING ............................................................................................................................................... 358
9.6. SIMIALR STUDIES ON THE CONVENTIONAL AND SELF-COMPACTING CONCRETE
SLABS 336
9.7. COMPARISON OF TIME-DEPENDENT DEFLECTION IN LIGHTWEIGHT, SELF-
COMPACTING AND CONVENTIOANL CONCRETE SLABS .......................................................... 363
9.8. TIME-DEPENDENT DEFLECTION IN CODES OF PRACTICE ........................................... 366
9.9. RATIO OF LONG-TERM TO INSTANTANEOUS DEFLECTION IN LIGHTWEIGHT, SELF-
COMPACTING AND CONVENTIONAL CONCRETE SLABS .......................................................... 368
9.10. DEFLECTION RATE IN LIGHTWEIGHT, SELF-COMPACTING AND CONVENTIONAL
CONCRETE SLABS ............................................................................................................................... 374
9.11. SUMMARY ................................................................................................................................ 375
REFERENCES ......................................................................................................................................... 377

Chapter 10: Finite element modeling of instantaneous and long-term deflection of light-
weight concrete slabs

10.1. INTRODUCTION ...................................................................................................................... 385


10.2. FINITE ELEMENT MODELS FOR REINFORCED CONCRETE STRUCTURES ................ 387
10.3. DAMAGE MODELLING IN CONCRETE ............................................................................... 387
10.4. REINFORCEMENT MODELLING ........................................................................................... 393
10.5. EFFECT OF ANALYSIS STEPS ............................................................................................... 394
10.6. CONVERGENCE METHODS IN NON-LINERA ANALYSIS ............................................... 395
10.7. ATENA SOFTWARE ................................................................................................................. 396
10.8. TECHNICAL SPECIFICATIONS OF ATENA-V5 ................................................................... 397

d
10.9. CONCRETE CONSTITUTIVE MODELS IN ATENA ............................................................. 399
10.10. MODELLING OF REINFOCEMENT IN ATENA.................................................................... 404
10.11. CRITERIA TO SELECT PROPER ELEMENTS IN ATENA ................................................... 404
10.12. CREEP AND SHRINKAGE ANALYSIS .................................................................................. 411
10.13. MODELIING OF SHRINKAGE BEHAVIOUR IN ATENA .................................................... 413
10.14. SELECTION OF COMPATIBLE ELEMENTS WITH EXPERIMENTAL DATA .................. 414
10.15. EFFECT OF MAXIMUM AGGREGATE SIZE AND AGGREGATE INTERLOCK .............. 415
10.16. EFFECT OF MESH SIZE ........................................................................................................... 416
10.17. WIDTH AND HEIGHT OF THE SELECTED ELEMENTS..................................................... 417
10.18. EFFECT OF LOADING PADS IN PROGRAMMING ............................................................. 417
10.19. ADVANTAGES OF SLABS SYMMETRY IN NUMERICAL ANALYSIS ............................ 418
10.20. ANALYSIS OF SLABS UNDER INSTANTANEOUS LOADING .......................................... 419
10.21. TIME-DEPENDENT ANALYSIS OF SLABS .......................................................................... 433
10.22. EFFECT OF THE IMPLEMENTED CREEP MODELS ON THE LONG-TERM DEFLECTION
OF SLABS ............................................................................................................................................... 447
10.23. SUMMARY ................................................................................................................................ 459
REFERENCES ......................................................................................................................................... 462

Chapter 11: Summary and conclusions

11.1. SUMMARY ................................................................................................................................ 464


11.2. CONCLUSIONS ......................................................................................................................... 466
11.3. RECOMMENDATIONS FOR FURTHER RESEARCH ........................................................... 480

APPENDIX A: Material test

APPENDIX A. ......................................................................................................................................... 482

APPENDIX B: Manufacturing the slabs and test setup for load-deflection test, instantaneous
deflection monitoring and long-term deflection monitoring

APPENDIX B. ......................................................................................................................................... 495

e
Chapter 1: Introduction

CHAPTER 1

INTRODUCTION
Chapter 1: Introduction

Table of Contents

1.1. BACKGROUND ............................................................................................................................ 1


1.2. STATE OF THE PROBLEM.......................................................................................................... 2
1.3. OBJECTIVES AND SCOPE OF THE RESEARCH ...................................................................... 4
1.4. METHODOLOGY .......................................................................................................................... 5
1.4.1. Experimental Program ............................................................................................................ 5
1.4.2. Numerical Investigation ......................................................................................................... 6
1.5. LAYOUT OF THE THESIS ........................................................................................................... 6
REFERENCES ............................................................................................................................................. 8
Chapter 1: Introduction

Chapter 1: Introduction

A general background on the topic and state of the problem is explained. The objectives and
expected results and contributions of the study in this area are presented. The research
methodology and the layout of the thesis in accordance with the literature review, experimental
program and the numerical investigation are described. To emphasise the topic’s importance, a
sample of the huge economic loss due to the large deflection of one-way slabs in real
construction projects is given.

1.1. BACKGROUND

Light-Weight Concrete (LWC) is used worldwide in the construction industry where savings
made in the weight of structures is a major factor. The most common applications of the
lightweight concrete are in floor systems, bridge decks, masonry and offshore structures. LWC
is made by replacing some or all of the normal weight aggregate with lightweight aggregate.
Often the coarse fraction is replaced with lightweight aggregate and the fines fraction is normal
weight sand. According to the lightweight concrete density and its compressive strength at age
28 days, lightweight concrete can be classified into two broad categories of structural and non-
structural concrete.

Lightweight structural concrete is typically made using expanded clay, shale or slate.
Lightweight aggregates such as crushed limestone, granite and quartz weigh less than normal
weight aggregates due to the porous cellular structure of the individual aggregate particles.

Compressive strength, elastic modulus, splitting tensile strength and other mechanical
properties of the lightweight concrete are significantly affected by the material properties of the

1
Chapter 1: Introduction

lightweight aggregate used. The lightweight aggregate itself must possess desirable properties
such as adequate compressive strength, porosity, abrasion resistance and good bonding with the
cement paste. For this reason, non-structural lightweight aggregate such as Perlite, Vermiculite,
and Styrofoam are not considered suitable for structural concrete, but rather are uses in concrete
meant for insulation or lightweight filler.

Developments in concrete technology provide engineers, suppliers and contractors with new
methods and approaches to engineering problems. Many of these developments are engineered
solutions to technical and commercial problems by either improvement of current practices or
overcoming limitations in the existing technology.

The long-term deflection of concrete slabs is often the governing design criteria to determine
the appropriate thickness to meet the required serviceability limit state. Serviceability limit state
places a limit on the maximum deflection and cracking allowed in the slab. If the structure does
not meet serviceability state requirements, it may be deemed not fit for the intended purpose.
Excessive concrete slab deflections would be deemed as a serviceability failure as the deflection
has adverse effects on the functionality of the other connected members, while not being in and
of itself a catastrophic structural failure. The negative impacts of excessive slab deflections on
the structure include jammed doors and windows, cracking of masonry and plaster partitions
and perceivable floor deflection.

An extreme example of a serviceability failure would be the infamous Silverton Centre


Office Building constructed in 1983 in Canberra. The flat slab concrete structure was evacuated
in 1989 and demolished in 1994 due to the excessive slab deflections and cracking. The
excessive deflections of the structure resulted in failure and subsequent leaking of the building’s
curtain walls. The structure’s value was estimated to be worth AUD$20m in 1994; the lost
revenue due to the vacancy of the building is estimated to be AUD$20m, and the legal costs of
all parties involved in the subsequent litigation are estimated to be over AUD$25m [1]. In
addition, a reliable and universally accepted procedure of designing the long-term deflection of
concrete slabs has not been developed yet.

1.2. STATE OF THE PROBLEM

Concrete is the world’s most broadly used construction material. However, despite all
benefits related to the use of concrete, the considerable self-weight of concrete compared to the
other construction materials and its workability problems limit its use in some structures.

2
Chapter 1: Introduction

Deflection is a common concern in almost all design codes and technical drafts in relation to
the time-dependent behaviour of the concrete structure. Short-term and long-term deflections
are regulated in serviceability limit states. Due to nonlinear nature of the concrete, prediction of
long-term deflection is complex. Also, the complicated behaviour of concrete in the presence of
reinforcement makes the problem hard to study. To better prediction of the deflection,
comprehensive information on time-dependent and nonlinear behaviour of material is required.

In general, there are big differences between the predicted values of deflection and the real
values measured in the concrete members under service loads especially in long-term deflection.
The codes of practice give an overall estimation of long-term deflection as a multiple of the
short-term or elastic deflection of concrete. Such estimation sometimes may lead to an entirely
wrong prediction of deflection and consequently unsafe and unprofitable concrete design and
construction.

There are different methods in the design codes to control deflection and to ensure the
serviceability adequacy of flexural reinforced concrete members over long-term period.
Limiting the minimum thickness of a concrete member is one of these simple methods of
deflection control. This limit is just an initial estimation of the expected deflection and is not
able to guarantee the expected accuracy under the different conditions of loading, environmental
effects and impacts of the mechanical properties of utilized materials.

One major problem with the explicit calculation of deflections, particularly long-term
deflection, is that the agreement between the calculated deflection and what actually occurs
should be checked. It is commonly found that the calculated and measured values differ to a
great extent. The deflection of concrete slabs involves numerous factors including the geometric
properties of the structure, concrete material properties and load history. Long-term deflections
are the result of increase in the short-term deflection over time due to concrete inelastic
shrinkage and creep strains. The accurate calculation of long-term deflection in concrete slabs is
crucial as it is often the governing design criteria to determine the slab thickness to meet the
required serviceability limit state.

Finite element programs are useful to analyse the non-linear behaviour of concrete slabs most
notably two-way slab systems. However, they lack the functionality to accurately model the
long-term deflections due to complexity of the factors involved. Currently, there is no
commonly accepted finite element software to calculate the long-term deflections in concrete
slabs. Without industry guidelines, there is an increased risk of structures failing to meet the
required serviceability limit state using finite element methods, the consequences of which can
be very severe as in the case of the Silverton Centre office building.

3
Chapter 1: Introduction

There exists a need for both theoretical and experimental research of the critical factors which
affect the time-dependent long-term deflection of different types of concrete structures under
real day to day service loads.

Most of the equations to predict the time-dependent behaviour in codes of practice are for
Conventional Concrete (CC), and there are limited numbers of relationships for Self-Compacted
Concrete (SCC) and LWC. SCC and LWC are results of recent technologies in the construction
industry. Hence, due to different mechanical properties of the used materials compared to the
conventional concrete, calculation and prediction of their short-term and long-term deflection by
utilizing the CC models and equations may not be justifiable.

Nejadi Gilbert (2005) [2] tested six one-way slabs made from CC under sustained loads for
about 400 days in the University of New South Wales (UNSW). Later, Aslani and Nejadi
(2014) [3] conducted a similar experiment on the identical slabs with similar loading, span
length and sections with different type of concrete. They used SCC to make the identical one-
way slabs.

Since LWC is very prevalent in the building construction in Australia and there are limited
studies on the long-term behaviour of this type of concrete, the current study examines the
identical LWC slabs with the previously tested CC and SCC slabs under long-term loading.
Therefore, to have a comparison with existing data of the previously conducted studies by
Nejadi and Gilbert (2005) [2] and Aslani and Nejadi (2014) [3], the same one-way slabs made
of LWC, with the similar loading, span length and sections have been investigated.

1.3. OBJECTIVES AND SCOPE OF THE RESEARCH

1- To monitor the affecting parameters associated with the fresh and hardened properties
of LWC;

2- To investigate and compare the fresh properties of CC, SCC and LWC (e.g. slump and
fresh concrete density);

3- To propose a new relationship between mixture design, slump and compressive strength
of SCC and LWC;

4- To study the hardened properties of LWC including the mechanical properties, bond
characteristics, shrinkage and creep to compare with those of CC and SCC;

5- To propose new mathematical models to predict the hardened properties of LWC


accurately;

4
Chapter 1: Introduction

6- To study the relationship between the compressive strength and modulus of elasticity of
LWC with the mixture design parameters by investigating the wide range of
experimental data of the utilized type lightweight concrete in the literature;

7- Comparison of the requirements of structural lightweight concrete in the fresh and


hardened state in different design codes and guidelines;

8- To obtain laboratory controlled data for LWC to calibrate, validate, and extend
analytical models that are being developed concurrently with the experimental program;

9- To study the mechanisms associated with the flexural deflection of slabs and influence
of the factors that affect their short-term and long-term deflections under sustained
service loads and proposing design guidelines for deflection calculations of LWC, SCC
and CC slabs;

10- Verification and validation of the analytical model by utilizing the finite element
program by implementing the proposed models of fresh and hardened LWC properties
into the ATENA software.

1.4. METHODOLOGY

This research comprises both experimental and analytical components. The experimental part
of the current study consists of laboratory investigation of the time-dependent flexural
deflection and monitoring of the strains in reinforced concrete elements of LWC. In addition,
the analytical part consists of two separate parts of verification of the recorded data of
experimental parts and the parametric study of the effective factors in time-dependent behaviour
and deflection of CC, SCC and LWC one-way slabs.

1.4.1. Experimental Program

Phase I: Testing of the fresh and hardened properties of LWC in the laboratory-controlled
conditions by utilizing the standard cylinders, prisms and cubes according to Australian
standards and test methods.

Phase II: Monitoring and recording the short-term (instantaneous) deflection of slab specimens.

Phase III: Monitoring and recording the long-term deflection of slab specimens.

5
Chapter 1: Introduction

Phase IV: Comparing the results in above mentioned phases with the results of former
investigations by Nejadi and Gilbert (2005) and Aslani and Nejadi (2014) on identical slabs
with different types of concrete.

1.4.2. Numerical Investigation

Phase I: Parametric study of the short-term and long-term deflection of the one-way simply
supported slabs

x For parametric study, recorded data from LWC tests together with the previously
recorded data of one-way slabs made from SCC and some other conventional concrete
data available in the literature are implemented.

Phase II: Verification of the proposed analytical model to calculate the LWC slabs deflection
by utilizing the recorded data of deflection and strain changes during the test period.

x Simulation of the slab specimens by ATENA software and utilisation of the proposed
material properties of LWC. ATENA Finite Element software is a powerful tool to
investigate the effects of different parameters on LWC slabs deflection accurately.

1.5. LAYOUT OF THE THESIS

This thesis is presented in eleven chapters and two appendices. The current chapter (Chapter
one) explains the state of the problem, objectives and methanoyl of the research program.

Chapter Two of this thesis provides a review of previous LWC studies related to history and
development, advantages and limitations, fresh properties, mechanical properties, bond
characteristics, shrinkage and creep, and available full-scale time-dependent LWC structure
investigations.

Chapter Three is a state-of-the-art review of the hardened concrete properties and available
empirical equations. Also, in Chapter Three, equations are developed for each of the hardened
concrete properties. Besides, to have a comprehensive comparison between the mechanical
properties of LWC with CC and SCC, the existing models of predicting the mechanical
properties of LWC, SCC and CC have been collected and compared.

Chapter Four is a review of the time-dependent behaviour of hardened concrete and available
empirical equations. Also, in Chapter Four, the existing models of shrinkage and creep and
bond-slip models in LWC, SCC and CC in the literature are investigated.

6
Chapter 1: Introduction

Chapter Five reviews the mechanism of flexural deflection in the reinforced concrete
structures. Additionally, the structural behaviour of concrete, reinforcement and the interaction
between concrete and reinforcement are broadly discussed and the existing models of tension
stiffening and the effective moment of inertia in literature are presented.

Chapter Six describes the experimental program for material properties of proposed LWC
mixtures. Also, the resultant mechanical properties of LWC have been compared with the fresh
and hardened properties of CC and SCC mixtures used in the experiments by Nejadi and Gilbert
(2005) and Aslani and Nejadi (2012) respectively.

Chapters Seven, Eight and Nine include the second phase of the experimental program.
Chapter Seven describes the experimental program for LWC short-term flexural deflection.
Also, a review of the similar experimental programs on CC slabs by Nejadi (2005) [2] and SCC
slabs by Aslani (2014) [3] is presented. Results of the current study on the short-term deflection
(load-deflection) of LWC slabs are compared with the corresponding experimental results on
the above-mentioned CC and SCC slabs.

Chapters Eight and Nine contain the explanations on the structures, testing setup, loading
magnitudes, monitoring system and the experimental results of the instantaneous and long-term
deflection of slabs, respectively. The experimental results of LWC slabs are compared with the
previously conducted experiments on the CC and SCC slabs. Additionally, new models are
proposed to calculate the instantaneous and long-term deflection of LWC, SCC and CC slabs.

Chapter Ten evaluates the finite element models simulated in the ATENA program. Also,
developed models in the previous Chapters about the LWC mechanical properties, bond
characteristics, shrinkage and creep are implemented in the ATENA program. At the end of this
Chapter, a parametric investigation using the proposed models to study the time-dependent
deflection has been performed and the results are summarised and discussed.

Finally, Chapter Eleven summarizes the conclusions drawn from the investigation of this
work and a proposal for future research is recommended.

Additional information including the raw data from creep and shrinkage tests, recorded
values by strain gauges embedded in one-way slabs and the deflection measured at two points of
each slab (mid-span and the border point of the high moment region of slab span) in LWC slabs
are provided in the Appendices.

7
Chapter 1: Introduction

REFERENCES

1. Gilbert, R., et al. The Risk and Responsibility of Structural Engineers–A Case Study of a
Structural Failure. in Proceedings of International Congress-Creating with Concrete. 1999.
2. Nejadi, S., Time-dependent cracking and crack control in reinforced concrete structures. 2005,
PhD thesis, The University of New South Wales Sydney, Australia.
3. Aslani, F., Experimental and numerical study of time-dependent behaviour of reinforced self-
compacting concrete slabs. PhD dissertation, University of Technology Sydney (UTS), 2014.

8
Chapter 2: Literature review

CHAPTER 2

LITERATURE REVIEW
Chapter 2: Literature review

Table of Contents

2.1. BACKGROUND AND DEVELOPMENT OF LIGHTWEIGHT CONCRETE ............................ 9


2.1.1. History .................................................................................................................................... 9
2.1.2. Development of lightweight concrete................................................................................... 12
2.2. MIX DESIGN OF LIGHTWEIGHT CONCRETE ....................................................................... 14
2.2.1. Mix design components and proportioning .......................................................................... 15
2.2.1.1. Monolithic lightweight concrete .................................................................................. 15
2.2.1.2. Aerated lightweight concrete ....................................................................................... 15
2.2.1.3. No-fine lightweight concrete ....................................................................................... 15
2.3. DEFINITION OF LIGHTWEIGHT CONCRETE ....................................................................... 16
2.3.1. ASTM-C330/M .................................................................................................................... 16
2.3.1.1. Standard Specification for lightweight aggregates in structural concrete .................... 16
2.3.1.2. Tests on Aggregates: ................................................................................................... 17
2.3.1.3. Tests on Concrete ........................................................................................................ 18
2.3.2. ACI-213R-2003 [1] .............................................................................................................. 18
2.3.3. ACI 211.2-98 [10] ............................................................................................................... 18
2.3.3.1. Weight method, specific gravity pycnometer .............................................................. 18
2.3.3.2. Volumetric method ...................................................................................................... 18
2.3.4. Experimental investigations ................................................................................................. 19
2.4. CLASSIFICATION OF LIGHTWEIGHT CONCRETE .............................................................. 19
2.5. ADVANTAGES AND DISADVANTAGES OF LIGHTWEIGHT CONCRETE ....................... 20
2.5.1. Advantages ........................................................................................................................... 20
2.5.2. Disadvantages....................................................................................................................... 22
2.6. FRESH PROPERTIES OF LIGHTWEIGHT CONCRETE ......................................................... 24
2.6.1. Workability........................................................................................................................... 24
2.6.2. Segregation ........................................................................................................................... 25
2.6.3. Testing of fresh properties .................................................................................................... 26
2.6.3.1. Slump ........................................................................................................................... 26
2.6.3.2. Air content ................................................................................................................... 27
2.7. MECHANICAL PROPERTIES OF LIGHTWEIGHT CONCRETE ........................................... 28
The present study investigates the time-dependent deflection of slabs made with LWC containing
Expanded Polystyrene (EPS) beads. This type of lightweight aggregate is commercially called BST
aggregates in Australia. ..................................................................................................................... 29
2.7.1. Compressive Strength........................................................................................................... 29
2.7.2. Modulus of Elasticity ........................................................................................................... 30
2.7.3. Tensile strength .................................................................................................................... 31
2.7.4. Modulus of rupture ............................................................................................................... 32
Chapter 2: Literature review

2.7.5. Bond characteristics of lightweight concrete ........................................................................ 33


2.8. FRACTURE ENERGY ................................................................................................................. 35
2.9. SHRINKAGE AND CREEP OF LIGHTWEIGHT CONCRETE ................................................ 37
2.9.1. Shrinkage of lightweight concrete ........................................................................................ 37
2.9.1.1. Drying shrinkage ......................................................................................................... 37
2.9.1.2. Autogeneous shrinkage ................................................................................................ 38
2.9.2. Creep of lightweight concrete .............................................................................................. 39
2.10. REAL-SCALE TIME-DEPENDENT DEFLECTION STUDY ............................................... 42
2.10.1. Birjandi and Clarck (1993), Deflection of lightweight aggregate concrete beams [76] ....... 42
2.10.1.1. Objectives .................................................................................................................... 42
2.10.1.2. Materials ...................................................................................................................... 43
2.10.1.3. Reinforcement ............................................................................................................. 43
2.10.1.4. Test arrangements and method .................................................................................... 43
2.10.1.5. Experimental results .................................................................................................... 44
2.11. SUMMARY.............................................................................................................................. 46
REFERENCES ........................................................................................................................................... 48
Chapter 2: Literature review

List of Figures
Figure 2.1: Weight spectrum of lightweight concrete (1lb/ft3=16.02 kg/m3) [4] ........................ 12
Figure 2.2: Typical ranges of densities of concretes made with various lightweight aggregates
(Approximate unit weight and use classification of light weight aggregate concrete) [4].......... 13
Figure 2.3: Natural and artificial lightweight aggregates used in lightweight concrete.............. 13
Figure 2.4: Test equipment and types of concrete slump............................................................ 26
Figure 2.5: Testing equipment of pressure and volumetric method to measure air content ....... 28
Figure 2.6: Classification of effective factors on modulus of elasticity of concrete ................... 30
Figure 2.7: Brazilian concrete splitting tensile test setup ........................................................... 31
Figure 2.8: ASTM-C78 beam specimen loaded at third-point loading [48]. .............................. 32
Figure 2.9: Schematic deformation of concrete in a tension specimen ....................................... 34
Figure 2.10: Effect of cracking in reinforced concrete member, A) a-portion of member; b-
bending moment distribution; c- bond stress distribution; d- concrete tensile stress distribution;
e- steel tensile stress distribution; f- flexural stiffness and B) a-forces; b-stress distribution; c-
strain distribution and d- bond variation between steel and concrete [51].................................. 34
Figure 2.11: Energy consumption for entire crack propagation ................................................. 36
Figure 2.12: Influence of the aggregates on the softening behaviour of lightweight concrete ... 37
Figure 2.13: Possible shrinkage curves for specimens of various concretes .............................. 38
Figure 2.14: Creep test set-up ..................................................................................................... 40
Figure 2.15: Various strains in concrete with time ..................................................................... 42
Figure 2.16: Test setup and loading arrangement of the LWC and CC beams [76] ................... 44
Figure 2.17: Deflection increase of LWC and CC beams with time [76] ................................... 46
Chapter 2: Literature review

List of Tables
Table 2.1: Specifications of lightweight aggregate for structural concrete in ASTM-C330/M.................. 16
Table 2.2: Grading requirements of lightweight aggregate for structural concrete in ASTM-C330/M ..... 17
Table 2.3: Concrete specimens containing lightweight aggregate in ASTM-C330/M ............................... 17
Table 2.4: Definition structural and non-structural lightweight concrete in codes of practice ................... 19
Table 2.5: Advantages of lightweight concrete .......................................................................................... 21
Table 2.6: Disadvantages of lightweight concrete application ................................................................... 23
Table 2.7: Relation between the workability of concrete and slump value ................................................ 26
Table 2.8: Allowed air content of lightweight concrete in ACI-211.2.98 .................................................. 27
Table 2.9: Compressive strength and modulus of rupture of conventional and lightweight concrete [50] 33
Table 2.10: Mixture proportions and the maximum aggregate size of beams ............................................ 43
Table 2.11: Mechanical properties of concrete and tensile bar ratio in LWC and CC beams .................... 44
Table 2.12: Recorded and calculated deflection of beams ......................................................................... 45
Table 2.13: Ratio of long-term to instantaneous deflection of LWC and CC beams ................................. 45
Chapter 2: Literature review

Chapter 2: Literature review

2.1. BACKGROUND AND DEVELOPMENT OF LIGHTWEIGHT


CONCRETE

2.1.1. History
The first known use of Light-Weight Concrete (LWC) dates back over 2000 years. There are
several LWC structures in the Mediterranean region, but the three most notable structures were
built during the early Roman Empire and included the Port of Cosa, the Pantheon Dome, and
the Coliseum. The Port of Cosa, built in about 273 B.C., used LWC made from natural volcanic
materials. The Pantheon, finished in 27 B.C., incorporates concrete varying in density from the
bottom to the top of the dome. The Coliseum, built from 75 to 80 A.D., is a gigantic
amphitheatre with a seating capacity of 50,000 spectators [1].

LWC has a surprisingly long history. It was first patented in 1923, mainly for use as an
insulation material. Although there is evidence that the Romans used air entertainers to decrease
density, this was not a real LWC.
Many different materials have been used in LWC since Roman days, and some have shown
remarkably better structural qualities than the Roman materials. Up to the 20th century, however,
such improvements were associated with the strength of the cementing materials used rather
than with improvements to the LWC.

The demand for LWC in modern construction applications, such as high-rise buildings,
offshore structures and long-span bridges, is increasing. This interest arises from the decreasing
amounts of many load-bearing elements, as well as the superior thermal properties of
lightweight compared to Conventional Concrete (CC) [2]. LWC can be obtained by introducing

9
Chapter 2: Literature review

gas (or foam) into the paste or by wholly or partially replacing the standard aggregate with low-
weight and preferentially low-cost components.

The implementation of LWC is usually necessitated by the reduction of the project cost,
enhanced functionality, or a combination of both. The first such improvement in the strength of
concrete came about in the course of Roman colonization when their need for widely scattered
building activity forced the use, on some occasions, of various impure grades of limestone for
the preparation of the required burned lime. In those instances where impurities were present in
the silica, alumina and iron oxides, the strength of concrete was found to be substantially greater
than where pure limes were used. This superior material was referred to as gray lime and
subsequently become known through Europe as Roman cement.

Light-Weight Aggregate and LWC are not new materials. LWC has been known since the
early days of the Roman Empire: both the Colosseum and the Pantheon were partly constructed
with materials that can be characterised as lightweight aggregate concrete (aggregates of
crushed lava, crushed brick and pumice). In the United States, over 100 World War II ships
were built in LWC, ranging in capacity from 3000 to 140000 tons and their successful
performance led, at that time, to an extended use of structural LWAC in buildings and bridges.

The logistic problems created by the entry of the United States into World War I in 1917
were compounded by the shortage of high-grade plate steel for building ships. The United States
Fleet Corporation, an arm of the Federal Government, was charged with planning a shipbuilding
program using materials other than steel. One of these materials was reinforced concrete, which
had already been used in shipbuilding in the Scandinavian countries.

The fundamental problem was the reduction of dead weight, and tests were made of concrete
made with natural aggregates, such as pumice, cinders and scoria, but these were found to be
unsuitable because of their comparatively low-strengths in addition to their permeability and
lack of uniformity.

The first commercial plant dedicated to expanded shale aggregate began operating in Kansas
City, Missouri, in 1920 under the name Haydite Company. Even so, there were few design
criteria available that could apply to applications of LWC in building construction, and little
inclination among architects, engineers and builders to risk their reputations by experimenting
with the new material. Two years later, a gymnasium addition to the Westport High School in
Kansas City was the first LWC building in history.

The first major project employing structural LWC was undertaken in 1928 and 1929, in the
form of an addition to the South-western Bell Telephone Company office in Kansas City. The
building was originally built as a fourteen-story structure, and the company had found that the

10
Chapter 2: Literature review

foundations and underpinning would support an additional eight floors, taking into account the
additional dead load of conventional heavyweight concrete.

The first structural LWC high-rise building was the Park Plaza Hotel in St. Louis (now the
Chase Park Plaza). Built in 1929, this twenty-eight-story structure made extensive use of
structure LWC in both frame and floor systems, as well as fire-proofing.

Among the more spectacular and sensational landmarks in the growth of LWC applications
have been its use in some major bridges, and in test buildings utilized in the historic atom bomb
tests at Yucca Flats, Nevada, in 1955. In the construction of the San Francisco-Oakland Bay
Bridge, for example, the use of LWC floor in the upper deck permitted weight reduction of 25.0
pounds per square foot or a total of 31.6 million pounds for the entire structure.

The term "Lightweight Aggregate" describes a range of particular use aggregates that have an
apparent specific gravity considerably below the normal sand and gravel which were at one time
used in almost all concrete.

According to Glenn A. Black (2004), the term "Lightweight Aggregate" describes a range of
special use aggregates that have an apparent specific gravity considerably below normal sand
and gravel which were at one time used in almost all concrete. So it is evident that the
lightweight aggregate is one of the critical elements that make concrete flexible and versatile
enough to make the overall structural design and specifications to meet the construction
requirements as argued by Leif Berntsson Satish Chandra (2002). It is also interesting to note
that the lightweight aggregate in the concrete that is made using light-weight materials also
provides a detectable level of compressibility as well as possessing the strength that can be
defined based on the composition thus making it a versatile and profitable process in production
process.

These lightweight aggregates will range from the extremely light materials used for insulative
and non-structural concrete all the way to expanded clays and shales used for structural
concrete. Since the lightness of these aggregates derives from the air trapped in each particle,
the more air that is trapped per particle unit, the lighter the weight and the better the insulation,
but, conversely, the lower the strength.

The range of lightweight aggregate is extensive, ranging from extremely light materials used
for insulation and non-structural concrete all the way to expanded clays and shales used for
structural concrete. This variety makes it clear that the lightweight aggregate in the concrete is
mainly aimed at achieving a high level of physical stability and compressibility through
effectively utilising the physical qualities of the aggregate materials. This is further justified by
the arguments of Leif Berntsson Satish Chandra (2002) who argues that the lightweight

11
Chapter 2: Literature review

aggregate in the concrete is a significant step towards innovation in the field of engineering
itself [3].

2.1.2. Development of Lightweight Concrete

Lightweight and free flowing (high flow-ability) concrete is a suitable material for a wide
range of purposes such as, but not limited to, panels and block production, floor and roof
screeds, wall casting, complete house casting, sound barrier walls, floating homes, void infill,
slope protection, outdoor furniture and many more applications.

The demand for lightweight concrete in modern construction applications, such as high-rise
buildings, offshore structures and long-span bridges is increasing. This interest arises from the
decreasing volume of many load-bearing elements, as well as the superior thermal properties of
lightweight concrete compared to conventional concrete.

Lighter concrete offers design flexibility and substantial cost-saving by providing less dead
load, improved seismic structural response, low heat conductivity and lower foundation cost
when applied to structures. In recent years, due to these advantages, there has been an interest in
production and investigation of the light or reduced-weight concrete.

The accompanying "Concrete Spectrum" in Figure (2.1) shows the relative weights and
differing applications of the various lightweight aggregates now in use. At the extreme left are
Vermiculite and Perlite, which are sometimes referred to as the "super lightweights." Concrete
can be made with these aggregates weighing as little as 240 or 320 kg/m3.

Figure 2.1: Weight spectrum of lightweight concrete (1lb/ft3=16.02 kg/m3) [4]

12
Chapter 2: Literature review

Air-dry densities of lightweight concrete ranging from 190 to 1900 kg/m3 are shown in
Figure (2.2) with different types of lightweight aggregates. Structural lightweight concrete at the
right end of the spectrum has strengths of 1800 kg/m3 and above [4].

Figure 2.2: Typical ranges of densities of concretes made with various lightweight aggregates (Approximate
unit weight and use classification of light weight aggregate concrete) [4]

In general, lightweight aggregates are broadly classified into two types: natural (e.g., pumice,
diatomite and volcanic cinder) and artificial (e.g., perlite, clay, sintered fly ash and expanded
shale). Lightweight aggregates have many sources. They may be produced using natural
materials such as shales, clays, pumice, diatomite, volcanic cinders, and slates. Artificial
materials (by-products) such as iron blast furnace slag, clay, sintered fly ash, and shale are other
sources of lightweight aggregates [4],[5],[6]. Figure (2.3) shows the sources and types of the
lightweight aggregate.

Lightweight aggregate

Natural lightweight aggregate Artificial lightweight aggregate

x Pumice x Artificial cinders


x Diatomite x Coke breeze
x Scoria x Foamed slag
x Volcanic cinders x Bloated clay
x Sawdust x Expanded shales and slate
x Rice husk x Sintered flay ash
x Exfoliated vermiculite
x Expanded perlite
x Thermo Cole beads
x Expanded polystyrene

Figure 2.3: Natural and artificial lightweight aggregates used in lightweight concrete

13
Chapter 2: Literature review

2.2. MIX DESIGN OF LIGHTWEIGHT CONCRETE

The main requirements of structural lightweight concrete from mix design are as follows:

- Compressive strength and density of the hardened concrete for the structural design;
- Consistency or workability for production purposes and entrained air volume.

Other properties may be required for a particular application such as frost resistance, water
impermeability, and the rate of chloride diffusion. For design work, formulas exist that are used
to calculate, for tensile strength or the modulus of elasticity, based on density and compressive
strength.

The aim of mix proportioning of concrete is to present a formula or a recipe according to


specifications. The procedure for proportioning is to combine different concrete ingredients to
attain the required properties of the fresh, as well as the hardened concrete. The properties of
concrete are chosen from the structural design and the requirements of safe and functional
structures. The requirements are mostly technical in nature, such as consistency, workability,
compressive strength, density, freeze-thaw resistance, water permeability, thermal conductivity,
and, not least, the service life of the structure. In many cases, there are requirements which are
implicit and expected of good quality concrete, i.e., stability against segregation of aggregate,
minor internal water bleeding, homogeneity after transportation, and compaction. Good
economy is the main driving force of all kinds of concrete structures.

LWC is quite similar to normal-weight (conventional) concrete, except that the whole or a
part of the aggregate consists of porous particles with the density being less than the natural
normal-weight aggregate of rock origin. In practice, the designation of LWC is classified
according to standardized strength and density classes. Structural LWC has a closed structure
which means that all voids between the aggregate particles are filled with a binder component or
cement paste[7]. The surfaces of the aggregate particles have to transfer the distributed stresses
from the surrounding binder. This is not the case in concrete with an open structure or no-fines
concrete where the internal stresses are concentrated at discrete contact points on aggregate
particles.

14
Chapter 2: Literature review

2.2.1. Mix Design Components and Proportioning

The common ingredients (mixture components) of LWC are as following:

x Binder, for example, type of cement and mineral admixture, i.e., silica fume, fly ash,
and slag;
x Fine and coarse aggregates, and their properties such as particle size distribution by
sieve analysis, bulk density, and particle density of the lightweight aggregate, water
content, and, eventually, water absorption;
x Chemical admixtures, such as plasticizers and air-entraining agents;
x Water.

In calculating the optimal solution for the strength and density of LWC at the start of the mix
design process, the results are dependent on the strength, density, and size of the available
lightweight aggregate. Sometimes, much is to be gained by changing one or two ingredients.
Based on the proportioning methods, most of the LWC can be designed for different
applications. There are many ways to produce lightweight concrete. Of these, three main types
can be identified as following:

2.2.1.1. Monolithic lightweight concrete


Monolithic lightweight concrete is a lightweight aggregate concrete with monolithic structure
in which the lightweight aggregate is used instead of normal-weight aggregate. The concrete
may be used as structural concrete and non-load-bearing concrete for heat insulation purposes.

2.2.1.2. Aerated lightweight concrete


This type of LWC is an aerated, foamed or gas concrete that relies on introducing large voids
within the concrete mass. There are examples where lightweight aggregate has also been added
to such mixes.

2.2.1.3. No-fine lightweight concrete


No-fines concrete as a low-density concrete is formed by omitting the fine aggregate, causing
a large number of interstitial voids. Coarse aggregate of normal density is used, but lightweight
aggregate gives considerably reduced weight and better heat insulation properties. The no-fines
concrete of lightweight aggregate is normally composed of the coarse lightweight aggregate of
fractions from 4 mm and upwards, for example, 4 to 8 or 10 mm. For no-fines, the fine
aggregate is omitted and the concrete consists of cement, water and coarse aggregate only. Quite

15
Chapter 2: Literature review

often, there is an addition of a mineral admixture such as fly ash or slag and a small addition of
filler or fine sand to the binder paste, mostly for economic and technical reasons [8]. No-fines
concrete is a lightweight aggregate in bulk with the particles surrounded by a coating of cement
paste or binder or mortar. The thickness may be about one millimetre and sometimes a little
more; for that reason, large voids exist between the aggregate particles. When graded aggregate
is used, a higher bulk density results in comparison to a one-size aggregate. The volume of
voids is decreased, resulting in higher strength and thermal conductivity.

2.3. DEFINITION OF LIGHTWEIGHT CONCRETE

There are various definitions and specifications for structural and non-structural LWC in
international codes of practice. Some of the most frequently used specifications of lightweight
aggregates in codes of practice are described below.

2.3.1. ASTM-C330/M

2.3.1.1. Standard Specification for lightweight aggregates in structural concrete


The main purpose for structural LWC in ASTM-C330/M (2014) [9] is reducing the density
while maintaining the compressive strength of the concrete. General specifications of
lightweight aggregate for structural concrete are defined in Table (2.1). The grading
requirements of allowed lightweight aggregates are presented in Table (2.2).

Table 2.1: Specifications of lightweight aggregate for structural concrete in ASTM-C330/M

Preparation method Physical Properties


Prepared by expanding, pelletizing, or
Clay lumps and friable particles less than
sintering products like blast-furnace slag,
Covered aggregate 2% by dry mass
clay, diatomite, fly ash, shale, or slate
types
Uniformity of grading change of fineness
Prepared by processing natural materials like
modulus in different samples should be
pumice, scoria, or tuff
less than 7%

16
Chapter 2: Literature review

Table 2.2: Grading requirements of lightweight aggregate for structural concrete in ASTM-C330/M

Percentage (mass) passing sieves having square openings


19.0 4.75 2.36 1.18 300 150 75
Nominal size 12.5 9.5
25.0mm mm mm mm mm μm μm μm
designation mm mm
(I in.) (3/4 (No. (No. (No. (No. (No. (No.
(1/2in.) (3/8in.)
in.) 4) 8) 16) 50) 100) 200)
Fine aggregate
Up to 4.75mm … … … 100 85-100 … 40-80 10-35 5-25 …

Coarse aggregate

4.75 to 25 mm 95-100 … 25-60 0-10 … … … … 0-10
10-50
4.75 to 19 mm 100 90-100 … 0-15 … … … … 0-10
40-80
4.75 to 12 mm … 100 90-100 0-20 0-10 … … … 0-10
80-100
2.36 to 9.5 mm … … 100 5-40 0-20 0-10 … … 0-10

Combined fine
and coarse aggregate
Up to 12.5 mm … 100 95-100 … 50-80 … … 5-20 2-15 0-10

Up to 9.5mm … … 100 90-100 65-90 35-65 … 10-25 5-15 0-10

Considering the mechanical properties, concrete specimens containing lightweight aggregate


under test should meet the requirements in Table (2.3).

Table 2.3: Concrete specimens containing lightweight aggregate in ASTM-C330/M

Average 28-day minimum Average 28-day


Calculated Equilibrium
Splitting Tensile Strength, minimum Compressive
Density max, kg/m3
, MPa Strength, MPa
1760 2.2 28
All lightweight aggregates 1680 2.1 21
1600 2.0 17
Combination of normal- 1840 2.3 28
Weight and lightweight 1760 21 2.1
aggregates 1680 2.1 17

The drying shrinkage of concrete specimens prepared, cured, and tested should not exceed
0.07%. Also, the following tests are required for aggregates and structural LWC:

2.3.1.2. Tests on Aggregates:


One representative sample is required for each test of organic impurities, staining, loss of
ignition, grading, bulk density, and clay lumps.

17
Chapter 2: Literature review

2.3.1.3. Tests on Concrete


At least three specimens are required for each of the following tests of concrete: compressive
strength, shrinkage, density, resistance to freezing and thawing, and the presence of pop-out
materials. At least eight concrete specimens are required for splitting tensile strength tests.

2.3.2. ACI-213R-2003 [1]

Structural aggregate meeting the requirements of ASTM-C330 for bulk density less than
1120 kg/m3 for fine aggregate and less than 880 kg/m3 for coarse aggregate is allowed in this
standard. The minimum required compressive strength at the age of 28 days is 17 MPa, by an
equilibrium density between 1120 and 1920 kg/m3. To achieve the requirements, application of
entirely lightweight aggregate or a combination of lightweight and normal-density aggregates is
allowed.

2.3.3. ACI 211.2-98 [10]

This standard provides applicable methods for selecting and adjusting mixture proportions for
structural lightweight concrete. The required specifications for structural LWC are as follows:

(a) Made with lightweight aggregates in accordance with ASTM-C330;

(b) Has a compressive strength in excess of 17 MPa at 28 days of age when tested in agreement
with methods stated in ASTM C330;

(c) Has an air dry weight not exceeding 1840 kg/m3 as determined by ASTM-C567.

In absence of prior information, the proportioning of first trial mixtures can be ascertained by
either of these methods:

2.3.3.1. Weight method, specific gravity pycnometer


This method applies to sand-lightweight concrete containing lightweight coarse aggregate
and normal weight fine aggregate. To estimate the weight of the LWC batch, the specific
gravity factor of lightweight coarse aggregate is determined and the first estimate of the weight
of fresh LWC can be made.

2.3.3.2. Volumetric method


This method is applicable for all lightweight and combinations of lightweight and normal
weight aggregates. The total volume of aggregates required is measured as the sum of the un-

18
Chapter 2: Literature review

combined volumes on a damp, loose basis. Of this amount, the loose volume of the fine
aggregate may make up 40 to 60 % of the total loose volume.

2.3.4. Experimental Investigations

Yang et al. (2014) [11] conducted a study to develop mixture-proportioning for structural
LWC to determine the unit content of each component to achieve the target values for the
slump, compressive strength, dry density, and air content of the concrete. Based on the
comprehensive database with 347 test results, which were compiled from concrete mixtures
with expanded fly ash or clay lightweight aggregates, they formulated the water to cement ratio,
and unit water content and provided a simple guide to proportion a trial mixture of structural
LWC. Because of the limited range and information of the dataset, the proposed approach was
optimized for the following requirements: the compressive strength range between 18 and 50
MPa, the range of dry density of concrete between 1200 kg/m3 and 2000 kg/m3, and the
maximum aggregate size being 19 mm or 25 mm.

2.4. CLASSIFICATION OF LIGHTWEIGHT CONCRETE

A classification of LWC can be based on the density of the concrete. There may be a relation
between the density and the compressive strength of lightweight aggregate concrete. A low-
density concrete has lower strength and a high-density concrete often has higher strength.
Structural lightweight aggregate concrete normally has compressive strength from at least 10 to
70 MPa.

The density and compressive strength of structural lightweight concrete are lower than those
for conventional concrete. According to Table (2.4), the definition of structural and non-
structural lightweight concrete, concerning density and compressive strength limits varies in
codes of practice and references.

Table 2.4: Definition structural and non-structural lightweight concrete in codes of practice

Reference Density (kg/m3) Compressive strength (MPa) Application


ACI 213R-03 [1] 800 to 2240 ൒ ͳ͹ Structural
ACI 211.2-98 [10] ൑ ͳͺͶͲ ൒ ͳ͹ Structural
ACI 213R(2014) [12] 1350 to 1900 ൒ ͳ͹ Structural
TS 2511 (1977) [13] ൑ ͳͻͲͲ ൒ ͳ͸ Structural
EN 206-1(2003) [14] 800 to 2000 8 to 80 Non structural
AS-3600 (2009) [15] 1800 to 2100 Structural

19
Chapter 2: Literature review

2.5. ADVANTAGES AND DISADVANTAGES OF LIGHTWEIGHT


CONCRETE

2.5.1. Advantages

The general advantages of lightweight concrete are its reduced mass and improved thermal
and sound insulation properties, while maintaining adequate strength. The reduced weight has
numerous advantages, not the least of them being a reduced demand for energy during
construction. In addition to lower density, better reinforcing steel-concrete bond, durability
performance, tensile strain capacity and fatigue resistance make the LWC preferable to normal
strength concrete.

Kivrak et al. (2006) describe the most common advantages of LWC compared to other types
of concrete as the following: [16]
x Reducing the seismic hazard for tall buildings and special concrete structures by
reducing the dead load of the structure;
x Reduction in the weight of the structure and consequently reducing the construction
costs;
x Facilitating the transport of concrete in the field of precast concrete structures;
x Better thermal insulation compared to ordinary concrete;
x Using artificial aggregates and eliminating the limitations in absence of natural
aggregates in all areas;
x High internal porosity which leads to maintaining a low apparent specific gravity.

Considering the massive consumption of natural raw materials in concrete industry,


Schackow et al. (2014) [17] recommended replacing the gravel in the concrete with expanded
polystyrene. Recycling of materials and waste reduction during the construction of buildings
and also in the recycling of demolition materials are the main reasons for their proposition.
They also mentioned the low thermal conductivity of LWC as an excellent option for
applications requiring good insolation. Lower dead load of structure, ease of cutting and
attaching filaments and unique sound isolation are other advantages of LWC in their study.

Since it is possible to achieve similar strength in LWC as conventional concrete, Rakoczy


and Andrzej (2014) [18] found LWC is a more efficient concrete in regard to better
strength/weight ratio than CC structural elements.

Lightweight concrete (LWC) is as a versatile material that has created keen interest and
considerable industrial demand in recent years in a wide range of construction projects. It has
been made lighter than conventional (normal weight aggregate) concrete [19], [20]. It has
obvious advantages of high strength to weight ratio, good tensile strength, low coefficient of
20
Chapter 2: Literature review

thermal expansion, and superior thermal and acoustic insulation characteristics due to the air
voids within its structure [21]. According to Newman et al. (2003) [22], it is possible to attain
the compressive strength over 60 MPa in LWC with oven dry density range of about 300 to not
exceed 2000 kg/m3 and thermal conductivities of 0.2 to 1.0 W/mK. These values can be
compared to those for normal weight concrete with approximately 2100–2500 kg/m3, greater
than 100 MPa and 1.6–1.9 W/mK.
Overall, the advantages of LWC can be summarized as the items in Table (2.5).

Table 2.5: Advantages of lightweight concrete


Advantage Description
Less mass is required to support additional weight. Structural reinforcement can be the less
Decreased dead Load
demanding
The design of economic concrete structures is accomplished by increasing the
strength/weight ratio. Structural LWC is typically 25 to 35 percent lighter and smaller than
Structure efficiency
conventional concrete. This translates into lighter superstructures and smaller loads for
substructure design.
Lower water Greatly reduced due to the diffusion of closed cells which prevents sponging. Also reduces
permeability problems caused by rusting rebar by eliminating the problem at its source
Higher Seismic In lower densities concrete can absorb shock. LWC is often used in ballistic tests because
resistance of this ability. Hammer blows can be absorbed without fracturing the concrete
Higher sound The transmission of sound is inversely related to the number of air/solid interfaces. LWC
absorption has a high number of these interfaces; thus, more sound is absorbed
Enhanced R-values, especially in the lower density range. Again, this is due to the
increased number of air/solid interfaces.
Greater Insulation
Structures can have much higher thermal efficiency resulting in lower heating or air
conditioning costs
Increased fire Greatly improved due to lower thermal conductivity. Spalling (scaling or flake chipping
resistance from heat) is reduced or eliminated
Lighter weight increases options for on-site casting. Forming can be swifter and easier due
Adaptability
to the less supported weight

Simplicity Ordinary tools can be used for alterations. It can be easily sewn and sculpted, and nailed or
screwed without pre-drilling
Handling capabilities are vastly improved. Concrete does not need to be cold, damp, dense,
Better handling
and hard to work with
Fewer failures at the pump from baling or compacting. Air bubbles from the foam act as
Pumping
miniature ball bearings in the mix. There are no settling out the problems
Having a greater yield/ton equates to reduced fuel consumption and lower transit costs for
Increased yield the producer. Adding foam up to 10% volume (and subtracting aggregate and water) will
not significantly reduce concrete strength - in some cases, it may improve
No Surface Bleed Reduction of excess water in the mix allows for sooner finishing. Flowability is achieved
water through the foam. Water can be utilized strictly for cement hydration

Placement Less handling weight makes all aspects of moving concrete easier, resulting in reduced
labour costs and quicker turnover

21
Chapter 2: Literature review

Table 2.5: Advantages of lightweight concrete (continued)

Advantage Description
Aesthetic The beauty of concrete is its ability to adapt to any shape, whether angular or curvilinear. If
improvements desired, it can be cast in different forms to make it look like wood
The problem of deforestation can be reduced by relinquishing the demand for timber and
substituting LWC in residential construction. Lightweight concrete in conjunction with
Environmental
tasteful design can be the solution for this environmental dilemma. Deforestation is also a
improvements
major contributor to global warming because vegetation removes a large quantity of carbon
dioxide from the atmosphere
One of the most extensive applications of structural LWC is in bridge re-decking where a
lower dead load is achieved. This often means that bridge widths, traffic lanes and the
Renovate and repair thickness of structural slabs can be increased while utilizing existing piers, footings and
other structural members. The use of LWC often allows the live load capacity of older
structures to be increased
LWC has been used in structures like bridge decks and building for over 50 years. The
Durability excellent in-service performance in these structures as well as in marine structures and
ships has demonstrated that lightweight concrete is a durable concrete.
Lightweight aggregate containing high internal moisture contents may be suitable for
conventional aggregates to provide internal curing. Higher cementitious concretes along
with a very low water to binder ratio are important parameters in the self-desiccation.
Improved hydration These concrete benefits significantly from the added internal moisture of properly pre-
wetted lightweight aggregates. Internal curing is particularly useful for concretes
containing high volumes of silica fume and other materials known to be sensitive to curing
procedures.

2.5.2. Disadvantages
Apart from advantages and benefits of application of LWC in the construction industry, it
also has several disadvantages. However, in many ways the benefits of lightweight concrete
overshadow its disadvantages.

Lightweight aggregate concrete often has a more complicated production process than
Conventional Concrete (CC) because the porous aggregates have excessively high water
absorption and low density. Water absorption by the aggregates results in severe slump loss and
a rapid setting time for the fresh concrete.

Sim et al. (2014) [23] compared LWC made with artificial lightweight aggregates and
conventional concrete and mentioned some weak points of LWC as following:
- Artificial lightweight aggregates used in LWC are commonly weaker and have
smoother surfaces than normal-weight aggregates. Hence, LWC has lower tensile
strength, lower cohesion between aggregate particles and pastes, and a more brittle

22
Chapter 2: Literature review

descending branch after peak stress compared with CC of the same compressive
strength;
- Cracks in LWC frequently propagate by passing through the aggregates; so the
interlocking capacity of aggregates along the crack panels will be significantly reduced.
Hence, the crack opening displacement and width of the fracture process zone can be
expected to be smaller in LWC than in conventional concrete.

Schackow et al. (2014) [17] used Expanded Polystyrene (EPS) to produce LWC, however
they mentioned that the increase in the content of EPS resulted in a significant reduction in the
compressive strength. Table (2.6) shows some other disadvantages of using LWC.

Table 2.6: Disadvantages of lightweight concrete application

Disadvantage Description
Sensitivity of mixture LWC is very sensitive to the water amount in the mixture
Longer mixing time Mixing time is longer than conventional concrete to assure a proper mixing
low compressive strength Compressive strength in LWC is less that of conventional concrete
The modulus of elasticity of the concrete with lightweight aggregates is about 0.5 to
Lower modulus of elasticity 0.75 to that of the normal concrete; hence, more deflection is there in LWC
members
Due to the higher moisture content of LWC, drying times are typically longer than
regular concrete to levels that might be considered adequate for application of floor
Longer drying time
covering materials. Typically, a 0.5 water to cement ratio slurry is used as a base
mixture for lightweight concrete.
Usually, the cover in LWC structure is 25mm more than that of normal concrete
Thicker reinforcement cover because of its increased permeability and also concrete carbonates rapidly by which
the protection to the steel by the alkaline lime is lost.
Pre-wetting aggregates require wetting before using, to achieve a high range of
saturation. Some concrete producers may not have the capability pre-wetting
Pre-wetting problems
lightweight aggregates in cold weather if temperature controlled storage is not
available
LWC is usually more expensive than ordinary concrete, and its production process
Expensive concrete
needs considerably more attention than that of the traditional concrete

LWC structure has a tendency to behave in a more brittle way than conventional
More brittle concrete concrete, special attention should be given to load-bearing mechanisms, where
spalling or splitting actions govern the behaviour

Cracks in LWC often spread by passing through the aggregates not the interlock of
Cracking of aggregates
the aggregates

23
Chapter 2: Literature review

2.6. FRESH PROPERTIES OF LIGHTWEIGHT CONCRETE

Although the general producing process of LWC and CC are similar in regard to mixture
components and curing conditions, there are considerable differences in properties of fresh
concrete between LWC and CC.

Main properties of fresh concrete are consistency, workability, segregation and bleeding. To
study these parameters, air content, fresh density, slump and setting time are investigated in
engineering and research applications of concrete.

Consistency or fluidity of concrete is a major component of workability and refers to the


wetness of the concrete. However, it must not be assumed that the wetter the mix the more
workable it is. If a mix is too wet, segregation may occur with resulting honeycombing,
excessive bleeding and sand streaking on the formed surface. On the other hand, if the fresh
concrete is too dry, it may be difficult to place and compact, and the segregation may occur
because of lack of cohesiveness and plasticity of the paste.

2.6.1. Workability

Workability is a term to describe the properties of fresh concrete and consists of the four
partial properties of concrete namely, mix-ability, transportability, mould-ability and
compatibility. Workability is often defined as the amount of mechanical work required for full
compaction of the concrete without segregation.

American Concrete Institute (ACI) defines the workability as the ease of placement of
concrete. This term is defined in ASTM-C125 as the property determining the effort required to
manipulate a freshly mixed quantity of concrete with minimum loss of homogeneity. The term
“manipulate” includes the operations of placing, compacting and finishing the concrete. Neville
and Brooks (2010) define the workability as the amount of useful internal work required to
produce full compaction [24].

Mehta and Monteiro (2006) indicate that workability is a composite property, with at least
two primary components: a) Consistency (or fluidity) which he describes as the “ease of flow”
of the fresh concrete, and b) Cohesiveness which he describes the ability of fresh concrete to
hold all ingredients together as the “tendency not to bleed or segregate”. Segregation risk
increases with increasing flowability which is generally higher in LWC with, respect to
conventional concrete. Hence, the control of workability and proper mix design is much more
critical for LWC than for conventionally placed concrete [25].

The followings are effective parameters in workability of fresh concrete:

24
Chapter 2: Literature review

- Water content in the concrete mix;

- Content and properties of cement;

- Shape, size, texture, maximum size and grading of the aggregates;

- Temperature and humidity of concrete mix;

- Compaction and placement method of fresh concrete.

Kupaei et al. (2013) [26] investigated the production of geopolymer LWC using two locally
available waste materials. They found that the fibres in the mixture absorb the existing water in
the paste and decrease the workability and also reduce the contact area between the lightweight
aggregate surface and the mortar, which causes a weak interfacial transition zone. To improve
the workability, they used Naphthalene sulphonated super plasticizer.

According to Alengaram et al. (2011) [27], with the same grade of concrete strength, LWC
by 20% lower density, shows higher workability, compared to CC.

2.6.2. Segregation

Having a broad conception of the effect of each component in mix design may improve the
capacity to reduce the unwanted effects. For example, good knowledge about the optimum
water to cement ratio brings down the segregation risk and enhances the flowability of the fresh
concrete. Segregation also makes the concrete weaker and less durable.

Expanded clay, granulated slag, perlite or vermiculite, and polymer materials are frequently
used as lightweight aggregate in LWC. Due to closed cavities, water absorption is high and it is
difficult to estimate the required water volume. Raising the water to the surface during the
mixing, accompanied by the tendency of lightweight aggregates to float up, increases the
segregation risk [28], [29].

Mazaheripour et al. (2011) [30] recommends applying the mix design method of high-
performance concrete for LWC to avoid the segregation problem and to keep the strength of
concrete high, despite applying it to light-weight aggregate.

25
Chapter 2: Literature review

2.6.3. Testing of Fresh Properties

2.6.3.1. Slump

The slump test measures the concrete flow. Higher slump concretes that are still cohesive
flow more readily and are easier to pump. However, excessively high slump concrete can
separate and cause plugs in the pump or the line.

The addition of a water reducing agent accompanied by a reduction in cement and water can
produce a mix that is hard to pump.

Slump and compacting factors can be reliable indicators of workability assessment. Table
(2.7) presents the relation between workability level and the range of slump in the concrete. The
Compacting Factor (CF) is defined as the weight ratio of partially-compacted to fully-
compacted concrete.

Table 2.7: Relation between the workability of concrete and slump value

Workability Slump (mm) Compacting factor Application


Very low 0-25 0.78 Roads-pavement
Low 25-50 0.85 Foundations
medium 25-100 0.92 Reinforced concrete
Reinforced concrete (highly
High 100-175 0.95
reinforced)

Measurement of slump in fresh concrete is typically performed using the method presented in
ASTM-C143 [31]. Australian code (AS-1012.3.1) [32] sets out the manner for determining the
slump of concrete when the nominal size of aggregate does not exceed 40 mm. Figure (2.4)
shows the equipment and types of the slump of concrete.

Figure 2.4: Test equipment and types of concrete slump

26
Chapter 2: Literature review

Dinakar (2013) [33] investigated the effect of water to binder ratio in regard to the
workability of conventional concrete and fly ash LWC. He found that for a given binder ratio
and the same super plasticizer dosage, the slump in LWC was recorded as slightly more than in
the CC. This observation was the same for all water/binder ratios.

2.6.3.2. Air content

The entrained air content (Vair) per unit volume of the fresh concrete is a major factor in fresh
concrete to achieve the target slump and mechanical properties. The entrained air in the LWC
mixture is designed to resist the freezing and thawing cycles. Improved workability of concrete
mixtures, especially those containing less water and cement, rough textured aggregates or
lightweight aggregates are some other benefits of air-entraining process. Hence, air entrainment
is a usual method of making mass concrete and lightweight concrete mixtures [17]. The air
voids in LWC also make it less vulnerable to stiffness and strength deterioration compared to
CC [34].

Moreno et al. (2014) investigated the effect of air voids (entrapped/entrained air) in
compressive strength of LWA and verified their experimental results by Equation (2.1) [35].

ࢌ࢘ࢋ࢒ ൌ ૚૙ି૜Ǥૡ૝ࢂࢇ࢏࢘ “ǤሺʹǤͳሻ

Where; frel is the relative compressive strength; and Vair is air-concrete ratio by volume.

The amount of entrained air recommended for LWC in ACI-211.2.98 [10] is limited as
shown in Table 2.8. According to ACI-213R-03 [1], the air content of LWC should not vary by
more than ±1%-1/2 % percentage points from the target value to avoid adverse effects on
compressive strength, concrete density, workability, and durability.

Table 2.8: Allowed air content of lightweight concrete in ACI-211.2.98

Conditions Air content (%) Max. aggr. size (mm)

LWC subjected to freezing and thawing or to 4-6 20


deicer salts 4.5-7.5 10

The measurement of air content in the fresh concrete of normal-density is typically performed
using the pressure method (ASTM-C-231) [36]. The volumetric method of measuring air, as
described in ASTM-C173 [37], is the most reliable method of measuring air in LWC. This
method relies on displacement of air with water in a vessel of pre-calibrated volume. Figure
27
Chapter 2: Literature review

(2.5) shows the testing equipment for Pressure method (left) and volumetric method to measure
the air content in fresh concrete.

Figure 2.5: Testing equipment of pressure and volumetric method to measure air content

Australian standard (AS-1012.4) [38] sets out a method for determining the air content of
freshly mixed concrete from observation of the change in volume of the concrete when it is
subjected to an increased air pressure. In the absence of testing facilities, this standard also gives
a simplified equation to find the air content of the concrete.

2.7. MECHANICAL PROPERTIES OF LIGHTWEIGHT CONCRETE

The early evaluation of hardened concrete properties is crucial and the compressive strength
of concrete is a fundamental parameter to estimate its other mechanical properties. However,
there is no direct relation to obtaining the compressive strength, and it has to be predicted by
experimental studies, destructive and non-destructive tests. Hardening properties of concrete
come directly from its fresh properties. The problem is that following the hardening process, the
quality and mechanical properties cannot improve. Structural behaviour of concrete relies on
mixing proportions and material properties of the composite system which cannot change after
hardening. Therefore, obtaining a relation to predict the hardening properties from the fresh
state and mixing proportions would be a great achievement in the concrete industry. However,
there are numerous models of international standards and codes of practice for the early
assessment of hardened properties of concrete at different ages. Sometimes these models give
under or overestimated predictions and strongly affect the targeted design and serviceability
aspects.

28
Chapter 2: Literature review

The present study investigates the time-dependent deflection of slabs made with LWC
containing Expanded Polystyrene (EPS) beads. This type of lightweight aggregate is
commercially called BST aggregates in Australia.

Despite considerable investigation of the mix proportioning and the relationship between
mechanical properties and fresh state properties of conventional concrete in the literature, there
are limited studies reported in the case of LWC, particularly LWC containing BST aggregates.

In this study, test results reported in the literature are used to evaluate the influence of LWC
mixture components and proportioning of the mechanical properties. The collected mechanical
properties will be compared with those of CC and Self-Compacting Concrete (SCC). The
existing models for mechanical properties of LWC, CC and SCC are presented and their
predictions for different properties will be compared. The applicability of the presented
prediction models are evaluated in Chapter Three. Other available prediction models for CC and
SCC are also assessed to check their capability of reasonably predicting the mechanical
properties of LWC. In conclusion, new models are proposed in this study to predict the
mechanical properties of LWC.

2.7.1. Compressive Strength

Compressive Strength (CS) gives an overall view of the quality of concrete. It is the most
important specification for quality characterization and concrete classification. Usually, the
other mechanical properties of concrete such as modulus of elasticity, stress-strain curve, tensile
strength, flexural strength (modulus of rupture), and bond strength are expressed in terms of the
compressive strength.

The compressive strength of structural LWC is specified according to the design


requirements of the structure. Normally, compressive strength of lightweight concrete is ranging
from 21 to 35 MPa; and less frequently up to 48 MPa or higher. Although some lightweight
aggregates are capable of producing very high strengths consistently, it should not be expected
that concrete made with every lightweight aggregate classified as “structural” can consistently
attain the higher strength values.

Kim et al. (2012) [39] compared the effect of mixture design on mechanical properties of air
entrained LWA and CC and reported the considerable influence of the entrained air content on
compressive strength and other mechanical properties of LWC. The air content in the LWC
matrix had a decreasing effect on compressive strength. The compressive strength increases
with matrix content, for the aggregate used and for air/matrix ratios of approximately 22% and
under; otherwise, compressive strength decreases.
29
Chapter 2: Literature review

2.7.2. Modulus of Elasticity

Modulus of Elasticity (MoE) is one of the most important characteristics of concrete that is
required in every design and evaluation process of concrete structures. MoE is also an essential
parameter in determination of the strain distributions and displacements, especially when the
elasticity considerations govern the concrete structure design.

Due to the brittle behaviour of concrete, various parameters affect MoE. Topcu and Ugurlu
(2007) [40] defined the effecting parameters as shown in Figure (2.6).

Figure 2.6: Classification of effective factors on modulus of elasticity of concrete

Two-phase (Mehta and Monterio (2006)) and three-phase (Nilsen and Monteir (1993))
models are the most frequent models for the researchers to explain the MoE of concrete. The
third phase known as Interfacial Transition Zone (ITZ), deals with the interface between mortar and
aggregate in concrete [35].

Nguyen et al. (2014) [41] studied the effect of mixture parameter and volume fraction on
thermal and mechanical properties of LWC. According to test results of seven types of fine and
coarse lightweight aggregates from three different environments in twelve mixtures, they
showed a linear decrease in the modulus of elasticity and compressive strength in reducing the
LWC density. Replacing normal-weight fine aggregate by lightweight fine aggregate by 100
kg/m3, decreases the compressive strength and the modulus of elasticity by 2.3-3.8 MPa and
1.7-2.6 GPa respectively.

Byrad et al. (2014) [42] compared the cracking tendency of LWC and CC in bridge decks.
They used expanded shale, clay, and slate lightweight coarse and fine aggregates to produce the
LWC. The results showed a decreasing tendency of MOE by increasing the amount of pre-
wetted lightweight aggregate in the concrete. The significantly reduced MOE and coefficient of

30
Chapter 2: Literature review

thermal expansion also led to a considerable overall delay in early age cracking in bridge deck
concrete.

2.7.3. Tensile Strength

Tensile strength of concrete is an effective parameter to evaluate the structural properties of


LWC such as the bond strength, shear resistance, anchorage and resistance to cracking. Despite
sufficient evidence about the effect of tensile capacity on total strength of concrete, it is usually
neglected in design and evaluation of reinforced concrete structures. Due to this tensile capacity,
concrete carries tensile strength between the cracks due to bond transfer between concrete and
reinforcement [43]. Three types of tests can determine concrete uniaxial tensile strength
including direct pull-out test, the modulus of rupture (flexural strength) test, and the splitting
tensile (Brazilian) test. Among the existing methods, the method illustrated in ASTM
C496/C496M-04 [44], is the most commonly used approach to determine the splitting tensile
strength. In this method, a concrete cylinder specimen is laid horizontally between the loading
platens of the testing machine and it is compressed along a vertical diameter. Figure (2.7) shows
the general arrangement of splitting tensile strength.

Figure 2.7: Brazilian concrete splitting tensile test setup

The lightweight aggregate content in the mixture influences the splitting tensile strength.
According to Gesoglu et al. (2004) [45], the splitting tensile strength in LWC gradually
decreases as the lightweight aggregate content increases from 30% to 60%. The maximum drop
in splitting tensile strength was in silica fume concrete with water/binder ratio equal to 0.32.

Tang et al. (2008) [46] compared the mechanical properties of high-strength sintered fly ash
LWC and crushed limestone CC and reported 25% lower splitting tensile strength in LWC.
From a review study on the Oil Palm Kernel Shell (OPKS) LWC, performed by Alengaram et
al. (2013) [47], inclusion of fly ash, silica fume and super plasticizer cause an improvement of
31
Chapter 2: Literature review

compressive strength from the typical range of 13-22 MPa to 37 MPa. Moreover, the
compressive strength reaches up to 48 MPa by adding crushed OPKS and limestone powder to
the LWC mixture. In this range, the splitting tensile strength was reported in the range of 6-10%
of the compressive strength.

Dinakar (2013) [33] used the same mixture proportions by volume for water/binder ratios
varying from 0.3 to 0.5 in LWC and CC mixtures and compared the mechanical properties . Use
of the natural coarse aggregate and Lytag coarse aggregates were the only differences in the
mixtures. As a result, they observed 7-9% variation of the ratio of splitting tensile/ compressive
strength in both LWC and CC specimens that was beyond the range given in European code of
practice, CEB-FIP.

2.7.4. Modulus of Rupture

In a reinforced concrete element subjected to flexural bending, immediately after the first
crack appears there is a sudden significant change in strain of tensile steel bars. Modulus of
rupture aims to accurately define of stress corresponding to the modification of concrete from
linear elastic condition to non-linear plastic condition. The maximum normal stress applied by
the ultimate bending moment under elastic behaviour assumption is defined as modulus of
rupture. The illustrated method in ASTM-C78/ C78M-10 [48] with the general arrangement of
the test specimen and flexural loading equipment shown in Figure (2.8) is used to define the
modulus of rupture in plain concrete.

Figure 2.8: ASTM-C78 beam specimen loaded at third-point loading [48].

Tassew and Lubell (2012) [49] investigated the mechanical properties of Light-Weight
Ceramic Concrete (LWCC) with a magnesium potassium phosphate binder. This binder was
applied to achieve the faster strength gain and lower the overall environmental impacts on the
concrete. Their experiment revealed that regardless of the aggregate type, the higher
32
Chapter 2: Literature review

water/binder ratio produces larger modulus of rupture in LWCC. Based on the existing model in
ACI-213(2003) [1] for Portland cement LWC, they also developed an empirical model with a
reduced coefficient for LWCC.

Alengaram et al. (2013) [47], concluded that in a high strength LWC made with Oil Palm
Kernel Shell (OPKS) LWC, the flexural tensile strength is 8-14% of the compressive strength.

Casanova-del-Angel and Vazquez-Ruiz (2012) [50] studied the production of LWC by


replacing the stone coarse aggregate with Poly-Ethylene Terephthalate (PET) lightweight
aggregate. Table (2.9) presents the recorded values of compressive strength and modulus of
rupture in CC and LWC. It is evident that despite the same range of compressive strength, the
modulus of rupture in LWC is lower than that of CC, particularly at higher strength classes.

Table 2.9: Compressive strength and modulus of rupture of conventional and lightweight concrete [50]

Conventional concrete Lightweight concrete


Compressive strength (MPa) 14.99 18.82 24.33 30.69 14.38 18.52 23.12 29.71
Modulus of rupture (MPa) 2.5 2.88 3.21 3.56 2.34 2.64 2.71 2.83

2.7.5. Bond Characteristics of Lightweight Concrete

Due to the presence of steel in reinforced concrete members, bar-concrete bond and
interaction consideration is inevitable. The concrete-reinforcement bond between two adjacent
cracks carries a certain amount of the tensile force normal to the cracked plane is illustrated by
tension stiffening that contributes to the overall stiffness of the member. Since bond stresses in
reinforced concrete members arise from the change in the steel force along the length, the effect
of the bond becomes more pronounced at end anchorages of reinforcing bars and in the vicinity
of the cracks. Figure (2.9) shows a schematic deformation of concrete in a tension region in the
vicinity of the deformed bar. Figures (2.10 A, B) show the idealized distribution of stress and
strain related to steel-concrete bond and the crack development in sections subjected to flexural
moment and direct tension respectively. It is, in fact, idealized distribution of stress and strain
between cracks of bond stress, concrete tensile stress and steel tensile stress respectively. As
shown in Figure (2.10 B-f), since concrete is carrying some tension between the cracks, the
flexural rigidity is clearly greater between the cracks, than at the cracks [51].

33
Chapter 2: Literature review

Figure 2.9: Schematic deformation of concrete in a tension specimen

A) Under flexural moment B) Under direct tension


Figure 2.10: Effect of cracking in reinforced concrete member, A) a-portion of member; b- bending moment
distribution; c- bond stress distribution; d- concrete tensile stress distribution; e- steel tensile stress
distribution; f- flexural stiffness and B) a-forces; b-stress distribution; c- strain distribution and d- bond
variation between steel and concrete [51]

From the tension stiffening concept, concrete does not crack suddenly and completely, but
undergoes progressive micro-cracking (strain softening). Due to the bond between steel and
concrete, the intact concrete between adjacent primary cracks carries considerable tensile force.
This bond significantly makes the stress-strain characteristics of tensile concrete in cracked
reinforced concrete structures different from the stress-strain characteristics of the plain
concrete.

The bond stress between steel and concrete is the combination of friction, adhesion and the
support of the ribs in deformed reinforcing steel. When adhesion fails, some relative movement
34
Chapter 2: Literature review

between steel and concrete and consequently cracking happen and the friction and steel ribs go
into action.

Chen et al. (2004) [52] compared the bond behaviour of LWC and CC in two groups of low
and high strength concrete. They presented the following results from the tests:
- Shear stirrups in the section improve the nominal bond strength in both LWC and CC
about 20%. However, its effect on critical bond strength is less than 20%;
- The bond failure surface through the lightweight aggregates particularly in low-strength
LWC, while in CC specimens most aggregates are not broken. For higher strength
levels over 40 MPa, the cracking surface is similar in both types of concrete;
- Corresponding bond strength in LWC is relatively lower than that for CC. However in
high-strength LWC, the bond strength is greater than that of CC, due to higher mortar
strength.

Sancak and Simsek (2001) [53] compared the steel-concrete bond in LWC produced by
pumice aggregate with conventional concrete. By replacing mineral additives and silica fume in
different ratios with Portland cement, they found the negative effect of silica fume on the bond
strength of LWC. Also, the bond strength of non-additive LWC in their experiment was 24%
lower than that of CC.

Alengaram et al (2010) [54] compared the bond properties of grade 30 LWC, namely Oil
Palm Kernel Shell Concrete (OPKSC) with CC of similar strength. The experimental
observation showed a weaker bond between OPKSC and cement matrix, compared to the bond
between crushed granite aggregate and cement matrix in CC. They also reported that the steel-
concrete bond stress of OPKSC was 86% of the same grade CC.

2.8. FRACTURE ENERGY

Fracture energy is defined as the energy required to open unit area of the crack surface
(complete separation of a concrete body) and is found to be increasing as we approach crack tip.
It is a material property; however there are some studies in the literature of investigations of the
size effect of structures on the fracture energy. Fracture energy can be expressed as the sum of
surface creation energy and surface separation energy. It is reprehensive for cracking resistance
and fracture toughness of concrete and therefore, is considered as a material property of
concrete [55].

According to the energy balance principle, the action of force P on its displacement will be
stored in two forms of energy; one part, as strain energy and the other part as used for crack

35
Chapter 2: Literature review

propagation. Equation (2.2) shows the relation between applied work and stored energy. Figure
(2.11) also displays the energy consumption for the entire cracking period.


ࡳࡲ࢏ ࡭࢏ ൌ ࢃ࢏ െ οࢁ࢏ ൌ ‫׬‬૙ ࢏ ࡼࢊࢾ െ οࢁ࢏ “ǤሺʹǤʹሻ

Where GFi is the average energy required for unit crack propagation from crack initiation
until any time; Ai is the newly formed fracture area for this period; Wi is the applied work of
external force P; Ui is the elastic strain energy, and δi is the crack opening displacement

Figure 2.11: Energy consumption for entire crack propagation

Although the “work-of-fracture” method recommended by RILEM [56] is dependent on the


size and shape of the structure, it is still the most commonly used method to measure the
fracture energy of concrete.

Kaplan (1961) initiated the study of fracture mechanics in concrete and Kesler (1971) and
Walsh (1972) reported the inadequacy of classical Linear Elastic Fracture Mechanics (LEFM)
for concrete structures. Hilleberg et al. (1976, 1985) made a major advance in the fracture
energy estimation and improved the developed the crack model (fictitious crack model) to
concrete. This model predicts the flexural failure of un-notched plain concrete beams. Peterson
(1981) refined and improved this model. Bazant (2002) presented a brief history of concrete
fracture models and studied the size effect on the fracture energy [57].

Sim et al. (2014) [23] compared the effect of aggregates and specimen size on the fracture
energy of LWC and CC beams. They tested four types of beam according to RILEM (1985)
[56]. As a result, the fracture energy in LWC was lower than that of CC. It was tending to
increase as the maximum size of aggregate increased up to 8 mm, beyond which it remains

36
Chapter 2: Literature review

constant. It is an indication of minor or no contribution of the aggregate interlock to the fracture


energy of LWC with the maximum aggregate size above 8 mm.

Dehn (2004) [58] investigated the fracture mechanics of 15 different LWC mixtures with dry
densities ranging from 1300 to 1800 kg/m3 and low to high-strength range between 20-100 MPa.
They found the effect of coarse aggregate on the maximum splitting tensile strength and fracture
energy. Also, the results showed that the fracture mechanism is not dependent to the type of
utilized lightweight aggregate. Since the brittleness of LWC increases with decreasing the grain
density of the coarse and fine aggregate, they compared conventional concrete by Sand Light-
Weight Aggregate Concrete (SLWAC) and All Light-Weight Aggregate Concrete (ALWAC).
Natural aggregates own a much higher resistance to cracking than porous and lightweight
aggregates; this is why ALWAC can transmit tensile stress over shorter cracks than SLWAC.
Figure (2.12) shows the effect of natural coarse aggregate, sand and lightweight aggregate on
the roughness, internal friction and cracking of the concrete.

Figure 2.12: Influence of the aggregates on the softening behaviour of lightweight concrete

2.9. SHRINKAGE AND CREEP OF LIGHTWEIGHT CONCRETE

2.9.1. Shrinkage of Lightweight Concrete

2.9.1.1. Drying shrinkage


Withdrawal of water from the concrete stored in the unsaturated air causes the drying
shrinkage. A part of this movement is irreversible and should be distinguished from the
reversible moisture movement caused by alternating storage under wet and dry conditions. In

37
Chapter 2: Literature review

this type of shrinkage, based on the thermodynamic equilibrium principle between moisture
evaporation and the environment, the shrinkage must stop after all the moisture has evaporated.

2.9.1.2. Autogeneous shrinkage


The chemical processes of Portland cement hydration leads to another type of shrinkage,
called the autogeneous shrinkage, which is observed in sealed specimens, i.e., at no moisture
loss. It is caused partly by chemical volume changes, but mainly by self-desiccation due to loss
of water consumed by the hydration reaction. This type of shrinkage is a limited part of the total
shrinkage and terminates once all the constituents have reacted. It amounts to only about 5% of
the drying shrinkage in normal concretes, which self-desiccate to about 97% pore humidity.
However, it can equal the drying shrinkage in modern high-strength concretes with very low
water-cement ratios, which may self-desiccate to as low as 75% humidity [59].

Figure (2.13) shows the possible shrinkage strain (εsh) curves for specimens of various
concrete types with identical size, shape, age and environment.

Figure 2.13: Possible shrinkage curves for specimens of various concretes

Considering the effect of mineralogical character of aggregate on shrinkage, Troxell et al.


(1958) [60] showed that the shrinkage of concretes containing limestone or quartz aggregate in
CC is only about half that of concrete containing sandstone aggregate.

Best and Polivka (1959) [61] conducted 18 months monitoring of shrinkage in expanded
shale LWC specimens having compressive strengths of 20 and 35 MPa. They found
considerably high drying shrinkage and deformation compared to the sand/gravel CC.

According to Neville (1995) [62], the lightweight aggregate results in higher shrinkage values
than normal-weight aggregate. However, Nielsen and Aitcin (1992) [63] found different results

38
Chapter 2: Literature review

as the drying shrinkage of LWC incorporating expanded shale was 30% to 50% lower than the
companion CC after 28 days of curing and 56 days of drying. For LWC with dense fine
aggregates, Newman (1993) [64] found similar shrinkage performance with the CC. They also
observed less shrinkage cracking in LWC due to the relief of restraint by creep and the
continuous supply of moisture from the pores of the lightweight aggregate. The experiment by
Kayali et al. (1999) [65] was in agreement with Neville (1995) [62]. They tested LWC made of
Lytag aggregates (sintered fly ash) and observed that the drying shrinkage (about 1000με) was
nearly twice that of CC after 400 days.

From a series of experimental tests on expanded shale LWC and CC, Reichard (1964) [66]
revealed that the average one-year shrinkages of the lightweight and normal-weight concrete
were 720 and 500 micro strain respectively.

Gesoglu et al. (2004) [45] investigated the free and restrained shrinkage behaviour of LWC
specimens with different ratios of water/cementitious material (w/c) and different volumes of
coarse aggregates in the mixture. The results showed a significantly poor shrinkage cracking of
LWC compared to CC. Also, the mixtures containing the higher volume of lightweight
aggregates showed higher shrinkage during the first 50 days. By comparing the effect of w/c
ratio, the concrete with lower w/c had higher shrinkage, in which the addition of silica fume
reduced the difference noticeably.

Shekarchi et al. (2012) [67] tried to develop a method of predicting the long-term shrinkage
from short-term measurements of four different mixtures of concrete. They reported the
inadequacy of the existing models for all types of concrete, emphasizing that individual
coefficients with the general form of Equation (2.3) are required to apply to each type of
concrete.


ࢿ࢙ࢎ ሺ࢚ሻ ൌ ࢻ࡯૚ ൫ࢌሺ࢚ሻ൯ ૛  Eq. (2. 3)

Where; α: coefficient of concrete type; f(t): time-function term; C1: modification factor and
C2 is used to modify the time-function term.

2.9.2. Creep of Lightweight Concrete

Under sustained load, concrete continues to deform. This phenomenon is called creep. Creep,
along with instantaneous (or elastic) strain, is defined as the strain difference between the
loaded and load-free specimens. Creep can be defined as the increase in strain under sustained
stress. The load-induced strain under conditions of no moisture exchange is known as basic

39
Chapter 2: Literature review

creep. However, if a specimen is also drying while under load, there is an increase in creep and
the extra creep is called drying creep.

Figure (2.14) shows the schematic set-up for creep test in laboratory condition.

Figure 2.14: Creep test set-up

The creep of concrete, and particularly its dependence on the age of loading and water
content, has proven to be a challenging problem. Bazant (2001) [68] introduced other effective
parameters in creep of concrete as followings:
- Plastic flow;
- Consolidation theory;
- Load-bearing hindered adsorbed water;
- Bond breakage in slip and its reformation;
- Nonlinear deformations and cracking as a contribution of the Pickett effect;
- Solidification theory for short-term aging;
- Micro-prestress of creep sites in cement gel microstructures.

Hatt (1907) discovered the creep phenomena in concrete. Despite considerable efforts and
achievement by the well-known experts such as Troxell et al. (1958) [60], Pickett, McHenry,
L’Hermite et al. (1965), Arutyunian, Aleksandrovskii (1959), Powers, Hansen and Mattock
(1966), Neville et al., (1983); the creep of concrete is not yet fully understood [69].
40
Chapter 2: Literature review

Best and Polivka (1959) [61] adjusted the creep relation proposed by Shank (1953), for LWC.
According to their findings, LWC exhibits lower modulus of elasticity than CC and therefore,
greater elastic strain under a given load.

From a series of experimental tests on CC and LWC samples containing expanded shale,
Reichard (1964) [66] revealed that in general, the CC exhibits less creep and shrinkage than the
LWC. However, some lightweight concretes exhibited less creep and shrinkage than some
normal-weight concretes. He loaded the LWC specimens at 28 hours and CC specimens at
different ages. The average adjusted creep of LWC was 0.145% of that at 365 days, while the
corresponding average creep for CC was 0.106%. Considering the cement content in the
mixture, he also found that in general, the amount of cement needed to develop a given strength
has increasing effect on the "creep plus shrinkage''.

Different lightweight aggregates show different creep behaviour. According to reported


results of Swamy and Ibrahim (1973) [70], Solite LWC shows the same order for creep as for
Lytag LWC made with rapid hardening cement, however, it was about 20-37% less than that for
creep of expanded clay (Aglite) LWC. They confirmed the results reported by Ross (1937) [71]
and Lorman (1940) [72] and found a linear relationship between creep and stress/strength ratio
for some value between 0.33 and 0.5.

Arnaouti and Sangakkara (1984) [73] performed an experiment to propose a model to predict
the creep and shrinkage of a large reinforced LWC structure from the shrinkage data of a small
plain LWC prism. Gamble and Parrott (1978) [74] concluded a linear relationship between
drying creep and concurrent shrinkage of LWC specimen. In their experiment, the basic creep
after 30 days was not significant.

The creep of LWC and CC is considered as an independent parameter of the stress history.
However, Tasdemir et al. (1988) [75] performed some studies to predict the creep strain under
various stress levels and found that the history of loading is affecting the creep strain and there
is a good logarithm correlation between the age of loading and the creep function.

Birjandi and Clarke (1993) [76] studied the long-term deflection behaviour of reinforced
LWC beams using two types of lightweight aggregate namely Lytag and Leca. They found that
parallel calculation for ±20% higher or lower shrinkage and creep coefficients produces only
±6% difference in the final deflection values of LWC beams.

Figure (2.15) shows the general presentation of possible time-dependent strains and their
general comparison.

41
Chapter 2: Literature review

Figure 2.15: Various strains in concrete with time

2.10. REAL-SCALE TIME-DEPENDENT DEFLECTION STUDY

Despite numerous examples in the literature of the long-term deflection of conventional


concrete structures, the investigation on the long-term deflection of LWC beams is very rare.
This section describes the summary of that experimental study in terms of material properties,
test methods, geometry of specimens and the obtained results.

To cover the general objective of this study, the time-dependent deflection on other types of
concrete are presented in Chapter Four, and the results are compared with the time-dependent
deflection of LWC slabs.

2.10.1. Birjandi and Clarck (1993), Deflection of Lightweight Aggregate Concrete


Beams [76]

2.10.1.1. Objectives

The main objectives of the study were:

- Long-term deflection of LWC beams over a period of six months;


- To cover different densities of LWC beams and comparing the results with CC beams;
- To compare the recorded deflection with estimated deflection by using the requirements
in BS 8110: Part 2 [77].

42
Chapter 2: Literature review

Specimens: A total of 12 beam specimens with the same length (3.0 m) and section
dimensions (100 mm wide and 200 mm deep) were made for two ratios of tensile reinforcement
and three types of concrete. The minimum concrete cover in all beams was 15 mm.

2.10.1.2. Materials

Concrete: Details of the LWC and CC mixture used in the study is given in Table (2.10). All
beams were cast in timber moulds, demoulded after 24 hours and then cured under a plastic
chamber until the loading age. Humidity and temperature of the lab during the test was kept at
40-60% and 20OC respectively.

Table 2.10: Mixture proportions and the maximum aggregate size of beams

Surface dry mixture proportions (kg/m3)


Mixture Coarse aggregate
Coarse aggr. sand Water OPC-42.5
LWC-1 Lytag - sintered PFA (3-12 mm) 637 789 241 277
LWC-2 Leca - expanded clay (3-10mm) 305 777 197 447
CC Normal aggregate (10mm) 930 910 200 270

2.10.1.3. Reinforcement
Tensile reinforcement: two arrangements of ribbed tensile reinforcement (each in six beams)
were used in the study. In the first group, 2N10 giving the reinforcement ratio of 0.9% and in
the second group, 2N8 giving the reinforcement ratio of 0.58% were used as tensile bars.

Compression reinforcement: In all beams, 2N6 ribbed bars were used in the compression
side.

Shear reinforcement: To resist the applied shear in all beams, R6@175 mm mild plain bars
were used.

2.10.1.4. Test arrangements and method


Each pair of beams of the same mixture were placed back-to-back (Figure 2.16) with their
tension sides facing each other and simply supported. They were then loaded centrally, using a
spring system to maintain the applied load as approximately constant. The sustained loading of
7 KN and 10 KN was provided by a hydraulic jack and was kept for duration six months.

43
Chapter 2: Literature review

Figure 2.16: Test setup and loading arrangement of the LWC and CC beams [76]

Deflection of each specimen (the pair of beams) was measured at 5 points (5 opposite points
on top and bottom beam) along the lengths including the mid-span. The total deflection of each
spot was measured between two opposite points on the beams on a vertical line. Thus, the
measured value was the sum of the deflection at two beams.

2.10.1.5. Experimental results

The measured mechanical properties of the mixtures are presented in Table (2.11). Table
(2.12) shows the recorded and estimated instantaneous deflection (Δins) and long-term deflection
(Δlong) after six months in the beams.

Table 2.11: Mechanical properties of concrete and tensile bar ratio in LWC and CC beams

Beam specimen
Property Curing
CC-a CC-b LWC-1a LWC-1b LWC-2a LWC-2b
Steel ratio % 0.58 0.9 0.9 0.9 0.58 0.58
Air 29.6 28.9 35.6 38.1 31.4 31.4
fcu (MPa)
water 31 27.8 33.1 35 32.7 31
Air 2.69 2.63 2.73 2.97 2.37 2.55
ft (MPa)
water 2.64 2.85 2.86 2.97 2.89 2.73
Air 2230 2210 1810 1810 1650 1630
Density (kg/m3)
water 2280 2270 1960 1970 1720 1700
Ec(GPa) 25.7 24.6 17.4 18.9 18.5 17.2
Estimated creep coefficient 1.8 1.8 1.3 1.3 1 1
Estimated shrinkage (με) 300 300 400 400 400 400
44
Chapter 2: Literature review

Table 2.12: Recorded and calculated deflection of beams

Beam specimen
Property
CC-a CC-b LWC-1a LWC-1b LWC-2a LWC-2b
Applied load (KN) 7.1 10 7.1 10 7.1 10
Experimental 9.7 14.3 9.5 15.4 14.6 21.6
Theoretical (measured Ec) 13.5 14.9 10.9 15.9 14.5 22.1
Δins
Theoretical (estimated Ec, BS-
13.5 14.7 11.4 16.8 16.2 23.9
8110) [77]
Experimental 20.8 24.1 17.8 25.3 21.9 30.4
Theoretical (measured Ec) 22.7 24.7 21.3 27.2 24.6 33.2
Δlong
Theoretical (estimated Ec, BS-
22.7 24.5 22.1 28.7 27 35.8
8110) [77]

The ratio of long-term deflection to instantaneous deflection (Δlong / Δins) for LWC and CC
beams are presented in Table (2.13). In average, the Δlong / Δins ratio in CC is higher than that of
LWC beams. In addition, this ratio is the smallest in Leca LWC beams. The theoretical methods
predict the deflection with ±20% accuracy. The ratio estimated for LWC by the British standard,
BS-8110 [77], particularly for the lower density Leca LWC is the smallest among all the beams.

Table 2.13: Ratio of long-term to instantaneous deflection of LWC and CC beams

Beam specimen
Property
CC-a CC-b LWC-1a LWC-1b LWC-2a LWC-2b
Actual 2.144 1.685 1.874 1.643 1.5 1.407
Theoretical (measured Ec) 1.68 1.658 1.954 1.71 1.697 1.5
Δlong / Δins
Theoretical (estimated Ec,
1.68 1.67 1.938 1.7 1.667 1.498
BS-8110) [77]

Figure (2.17) compares the time-dependent deflection of Leca LWC and Lytag LWC with
CC.

45
Chapter 2: Literature review

Figure 2.17: Deflection increase of LWC and CC beams with time [76]

2.11. SUMMARY

A brief history of the invention and development of lightweight concrete, together with the
current classification of lightweight concrete in codes of practice are given. Also, the definitions
and the minimum requirements of structural and non-structural lightweight concrete in the
frequently used design codes such as ASTM-C330, ACI-211.2-98, ACI.213R-03, ACI.213R-14,
EN-206.1-03, TS-2511-97 and AS-3600-09 are given.

Investigations of performance and economic aspects of LWC are the concerns in producing
LWC with new types of natural and artificial lightweight aggregates. Various sources of natural
or artificial lightweight aggregates used to produce the lightweight concrete and mixing
methods are explained broadly. Advantages and disadvantages of the lightweight concrete
regarding the economic issues, mixing and preparation, technical performance, fire protection,
seismic load bearing, and some other effective parameters in the selection of the proper concrete
type for construction projects are discussed.
46
Chapter 2: Literature review

The definitions for structural and non-structural LWC in some international standards are
compared. Furthermore, the criteria for choosing the proper aggregates in LWC are discussed.
Fresh and hardened state characteristics of LWC are studied and the common methods of testing
and instruction are given in the first part of this chapter. Fresh LWC density, air entrained
content, workability, segregation and slump are some fresh state properties that are discussed
and compared with those for Conventional Concrete (CC). According to the investigated case
studies, LWC owns better workability than CC, however due to presence of lightweight
aggregates; there is a risk of segregation in LWC.

A review of the properties of fresh and hardened lightweight concrete and the relevant testing
methods are given. The workability, segregation, slump and air content are the topics related to
the fresh lightweight concrete. Also, the compressive strength, modulus of elasticity, tensile
strength, modulus of rupture, bond characteristics, and fracture energy are among the most
important mechanical properties of lightweight concrete which are discussed and compared to
the corresponding properties of the conventional concrete. The compressive strength and
density of LWC are less than those of CC in general. Besides, the mechanical properties of
LWC in the same range of compressive strength in conventional concrete are of lower values in
LWC.

Considering the time-dependent properties, the creep and shrinkage in lightweight concrete
are compared with those of the conventional concrete. In general, the shrinkage and creep
effects are more considerable in LWC due to its higher water absorption and reduced elasticity
in comparison to CC.

Despite the limited number of long-term experimental studies on the mechanical and time-
dependent properties of lightweight concrete structures in the literature, collection of all the
relevant experimental investigation as a means of strengthening the basis of the current research
has been attempted. Therefore, the most relevant experimental studies and the obtained results
are discussed. An intention of this study was to investigate the instantaneous and long-term
deflection of LWC slabs, however after a comprehensive review, only one example of related
case study was found on LWC beams in literature. Some highlights of that experiment in terms
of test setup, materials and specimens and the obtained results are presented and evaluated.
There are more investigations of other types of concrete in the literature that will be presented
and compared in Chapter Four of this study.

47
Chapter 2: Literature review

REFERENCES

1. ACI-213R-03, Guide for Structural Lightweight-Aggregate Concrete. ACI Committee 213,


American Concrete Institute, Farmington Hills, Mich. 48333-9094, 2003.
2. Vakhshouri, B. and S. Nejadi, Mix design of light-weight self-compacting concrete. Case Studies
in Construction Materials, 2016. 4: p. 1-14.
3. Nicolas Ali Libre, et al., Mechanical properties of hybrid fiber reinforced lightweight aggregate
concrete made with natural pumice. Construction and Building Materials 2011. 25: p. 2458–
2464.
4. A.M, N., Properties of Concrete. Pitman Publishing , 3rd Edition, 1981.
5. ACI-213R-87, Guide for Structural Lightweight Aggregate Concrete. ACI Committee 213,
American Concrete Institute, Farmington Hills, Mich. 48333-9094, 1999.
6. Miled K., et al., Compressive behavior of an idealized EPS lightweight concrete: size effects and
failure mode. Mechanics of materials, 2004. 36(11): p. 1031–1046.
7. Vakhshouri, B. and S. Nejadi, Self-compacting light-weight concrete; mix design and
proportions. Structural Engineering and Mechanics, 2016. 58(1): p. 143-161.
8. Chandra Satish and B. Leif, Lightweight aggregate concrete, science, technology and
application. William Andrew Publishing/Noyes, 2002.
9. ASTM-C330/C330M-1, Standard Specification for Lightweight Aggregates for Structural
Concrete. American Socity of Testing and materials (ASTM), 2014.
10. ACI-211.2-98, Standard Practice for Selecting Proportions for Structural Lightweight Concrete.
ACI Committee 211, American Concrete Institute, Farmington Hills, Mich, 1998.
11. Yang, K.-H., G.-H. Kim, and Y.-H. Choi, An initial trial mixture proportioning procedure for
structural lightweight aggregate concrete. Construction and Building Materials, 2014. 55: p.
431-439.
12. ACI-213R-03, Guide for Structural Lightweight-Aggregate Concrete. ACI Committee 213,
American Concrete Institute, Farmington Hills, Mich. 48333-9094, 2003.
13. TS-2511, Mix Design for Structural Lightweight Aggregate Concrete. Turkish Standards
Institution (TSI), Turkey, Ankara (in Turkish), 1977.
14. EN-206-1, Specification, performance, production and conformity. European Standard, 2000: p.
72.
15. AS-3600-09, Concrete structures. Australian Standard, 2009.
16. S. Kivrak, M.T., M. I. Onur, G. Arslan and O. Arioz, An economic perspective of advantages of
using light weight concrete in construction. 31st Conference on OUR WORLD IN CONCRETE
& STRUCTURES: 16 - 17 August 2006, Singapore, 2006.
17. Schackow, A., and, et al., Mechanical and thermal properties of lightweight concretes with
vermiculite and EPS using air-entraining agent. Construction and Building Materials, 2014. 57:
p. 190-197.
18. M., R.A. and A.S. Nowak, Resistance Factors for Lightweight Concrete Members. ACI
Structural Journal, 2014. 111(1).
19. El Zareef, M.A., Conceptual and structural design of buildings made of lightweight and infra
lightweight concrete. PhD thesis, Berlin University of Technology, 2010.
20. Babu, D.S., Mechanical and deformational properties, and shrinkage cracking behavior of
lightweight concretes. PhD thesis, national university of Singapore, 2008.
21. Abdulkareem, O.A., et al., Effects of elevated temperatures on the thermal behavior and
mechanical performance of fly ash geopolymer paste, mortar and lightweight concrete.
Construction and Building Materials, 2014. 50: p. 377-387.

48
Chapter 2: Literature review

22. Newman, J., Choo, B. S. and Owens, P, Advanced Concrete Technology Processes. Elsevier Ltd,
2003.
23. Sim, J., Yang, K., Lee, E., and Yi, S, Effects of Aggregate and Specimen Sizes on Lightweight
Concrete Fracture Energy. J. Materials in Civil Engineering, 2014. 26(5): p. 845-854.
24. Koehler E.P. and F.W. D., ICAR Project 108: Aggregates in Self-Consolidating Concrete.
Aggregates Foundation for Technology, Research, and Education (AFTRE). International Center
for Aggregates Research (ICAR), The University of Texas at Austin, 2007.
25. Neshat, M., et al., A Comparative Study on ANFIS and Fuzzy Expert System Models for
Concrete Mix Design. IJCSI International Journal of Computer Science Issues, 2011. 8(3).
26. Hosseini Kupaei, R., et al., Mix design for fly ash based oil palm shell geopolymer lightweight
concrete. Construction and Building Materials, 2013. 43: p. 490-496.
27. Alengaram, U.J.a., H. Mahmud, and M.Z. Jumaat, Enhancement and prediction of modulus of
elasticity of palm kernel shell concrete. Materials & Design, 2011. 32(4): p. 2143-2148.
28. Barrios, F.A., Acoustic emission techniques and cyclic load testing for integrity evaluation of
self-compacting normal and self-compacting lightweight prestressed concrete girders. 2010.
29. Juradin, S., G. Baloević, and A. Harapin, Experimental Testing of the Effects of Fine Particles
on the Properties of the Self-Compacting Lightweight Concrete. Advances in Materials Science
and Engineering, 2012. 2012: p. 1-8.
30. Mazaheripour, H., et al., The effect of polypropylene fibers on the properties of fresh and
hardened lightweight self-compacting concrete. Construction and Building Materials, 2011.
25(1): p. 351-358.
31. 143, A.-C., Standard test method for slump of hydraulic-cement concrete. aMERICAN Society
for Testing Materials (ASTM), West Conshohocken, 1997.
32. AS-1012.13, Methods of testing concrete, method 13-determination of the drying shrinkage of
concrete for samples prepared in the field or in the laboratory. Australian standard, 2014.
33. Dinakar, P. Properties of fly-ash lightweight aggregate concretes. Proceedings of the ICE -
Construction Materials, 2013. 166, 133-140.
34. Zanjani Zadeh, V. and C.P. Bobko, Nanomechanical characteristics of lightweight aggregate
concrete containing supplementary cementitious materials exposed to elevated temperature.
Construction and Building Materials, 2014. 51: p. 198-206.
35. Moreno, D., P. Martinez, and M. Lopez, Practical Approach for Assessing Lightweight
Aggregate Potential for Concrete Performance. ACI Materials Journal, 2014. 111(2): p. 123-
132.
36. ASTM-C231, Standard Test Method for Air Content of Freshly Mixed Concrete by the Pressure
Method. American Society for Testing Materials (ASTM), 2014.
37. C173, A., Standard Test Method for Air Content of Freshly Mixed Concrete by the Volumetric
Method. American Society for Testing Materials (ASTM), 2007: p. 8.
38. AS-1012.4, methods of testing concrete, method 4.1-determination of air content of freshly
mixed concrete- measuring reduction in concrete volume within increased air pressure.
Australian standard, 2014.
39. Kim, Y.-J., et al., Mechanical properties of aerated lightweight aggregate concrete. Magazine of
Concrete Research, 2012. 64(3): p. 189-199.
40. İlker Bekir Topcu, A.U., Elasticity Theory of Concrete and Prediction of Static E-Modulus for
Dam Concrete Using Composite Models. Digest, 2007: p. 1115-1127.
41. L.H. Nguyen, et al., Influence of the volume fraction and the nature of fine lightweight
aggregates on the thermal and mechanical properties of structural concrete. Construction and
Building Materials, 2014. 51: p. 121-132.
42. Byard, B.E.a., A.K. Schindler, and R.W. Barnes, Cracking Tendency of Lightweight Aggregate
Bridge Deck Concrete. ACI Materials Journal, 2014. 111(2).

49
Chapter 2: Literature review

43. Ian Gilbert, R., Tension Stiffening in Lightly Reinforced Concrete Slabs. Journal of Structural
Engineering, 2007. 133(6): p. 899-903.
44. C496/C496M-04, A., Standard Test Method for Splitting Tensile Strength of Cylindrical
Concrete Specimens. aMERICAN Society for Testing Materials (ASTM), West Conshohocken,
2004: p. 5.
45. Gesoğlu, M., T. Özturan, and E. Güneyisi, Shrinkage cracking of lightweight concrete made with
cold-bonded fly ash aggregates. Cement and Concrete Research, 2004. 34(7): p. 1121-1130.
46. Tang, W.C., T.Y. Lo, and W.K. Chan Fracture properties of normal and lightweight high-
strength concrete. Magazine of Concrete Research, 2008. 60, 237-244.
47. Alengaram, U.J., B.A.A. Muhit, and M.Z.b. Jumaat, Utilization of oil palm kernel shell as
lightweight aggregate in concrete – A review. Construction and Building Materials, 2013. 38: p.
161-172.
48. C78/C78M-10, A., Standard Test Method for Flexural Strength of Concrete (Using Simple Beam
with Third-Point Loading). aMERICAN Society for Testing Materials (ASTM), West
Conshohocken, 2010: p. 4.
49. Tassew, S. and A. Lubell, Mechanical properties of lightweight ceramic concrete. Materials and
Structures, 2012. 45(4): p. 561-574.
50. Casanova-del-Angel, F. and J.L. Vázquez-Ruiz, Manufacturing Light Concrete with PET
Aggregate. ISRN Civil Engineering, 2012. 2012.
51. Park, R. and T. Paulay, Reinforced concrete structures. 1975: John Wiley & Sons.
52. Chen HJ, Huang CH, and K. ZY, Experimental investigation on steel–concrete bond in
lightweight and normal weight concrete. Structural Engineering and Mechanics 2004: p. 141-
152.
53. Sancak, E., O. Simsek, and A.C. Apay, A comparative study on the bond performance between
rebar and structural lightweight pumice concrete with/without admixture. Int J Phys Sci, 2001.
6(14): p. 3437-3454.
54. U. Johnson Alengaram, H.M.a.M.Z.J., Comparison of mechanical and bond properties of oil
palm kernel shell concrete with normal weight concrete. International Journal of the Physical
Sciences 2010. 5(8): p. 1231-1239.
55. Cifuentes, H., M. Alcalde, and F. Medina, Measuring the SizeǦIndependent Fracture Energy of
Concrete. Strain, 2013. 49(1): p. 54-59.
56. TCM-85., R., Determination of the fracture energy of mortar and concrete by means of three-
point bend tests on notched beams. Materials and Structures, 1985. 18: p. 287-290.
57. Bazant, Z.P., Concrete fracture models: testing and practice. Engineering fracture mechanics
2002. 69: p. 165-205.
58. F, D., Fracture mechanical behaviour of lightweight aggregate concrete. In: Proceedings of the
5th International Conference on Fracture Mechanics of Concrete and Concrete Structures
(FraMCos-5), 2004. 2: p. 1097-1104.
59. Bazant, Z.P., Criteria for rational prediction of creep and shrinkage of concrete. ACI SPECIAL
PUBLICATIONS, 2000. 194: p. 237-260.
60. Troxell, G.E., J.M. Raphael, and R.E. Davis, Long-time creep and shrinkage tests of plain and
reinforced concrete. American Society for Testing Materials (ASTM), 1958. 91.
61. Best, C.H., and Polivka, M. , Creep of lightweight concrete. Magazine of Concrete Research,
1959. 11(33).
62. Neville, A.M., Properties of Concrete. 4th ed , Birmingham, UK, 2010.
63. Nilsen, U. and P. Aitcin, Properties of high-strength concrete containing light-, normal-, and
heavyweight aggregate. Cement, Concrete and Aggregates, 1992. 14(1).
64. Newman, J.B., Properties of structural lightweight aggregate concrete. in: J.L. Clarke (Ed.),
Structural Lightweight Aggregate Concrete, Chapman & Hall, London, 1993: p. 19-44.
50
Chapter 2: Literature review

65. Kayali, O., M.N. Haque, and B. Zhu, Drying shrinkage of fibre-reinforced lightweight aggregate
concrete containing fly ash. Cement and Concrete Research, 1999. 29(11): p. 1835-1840.
66. Reichard, T.W., Creep and drying shrinkage of lightweight and normal weight concretes.
National Bureau of Standards, Washington, D.C., Monograph 74, 1964.
67. Shekarchi, M., F. Ghasemzadeh, and S. Sajedi, Inverse analysis method for concrete shrinkage
prediction from short-term tests. ACI Materials Journal, 2012. 109(3).
68. Bazant, Z.P., Prediction of concrete creep and shrinkage: past, present and future. Nuclear
Engineering and Design 2001. 203: p. 27-38.
69. TC-69, R., State of the art in mathematical modeling of creep and shrinkage of concrete. In:
Bazˇant, Z.P., (Ed.), Mathematical Modeling of Creep and Shrinkage of Concrete, Wiley,
Chichester and New York, pp. 57–215; and in prelim. form: State-of-art report on creep and
shrinkage of concrete: mathematical modeling,. Fourth RILEM International Conference on
Creep and Shrinkage of Concrete, 1986, 1988: p. 41-80.
70. N., S.R. and A. K.L., Shrinkage and creep properties of high-strength structural concrete. Civil
Engineering and Public Works Review, 1973. 68(807): p. 855-868.
71. D., R.A., Concrete creep data. Structural Engineering, 1937. 15: p. 314-326.
72. Lorman, W.R., The theory of concrete creep. American Society for Testing Materials (ASTM),
1940. 40: p. 1082-1120.
73. Arnaouti, C. and S.R. Sangakkara, Creep and shrinkage ain lightweight-aggregate concrete.
Magazine of Concrete Research, 1984. 45(162): p. 43-49.
74. Gamble, B.R. and L.J. Parrott, Creep of concrete in compression during drying and wetting.
Magazine of Concrete Research, 1978. 30(104): p. 129-138.
75. Tasdemir, M.A., S. Akyuz, and N. Uzunhasanoglu, Creep of lightweight aggregate concrete
under variable stresses. Cement, Concrete and Aggregates, 1988. 10(2): p. 61-67.
76. Birjandi, F. and J. Clarke, Deflection of lightweight aggregate concrete beams. Magazine of
Concrete Research, 1993. 45(162): p. 43-49.
77. BS-8110, Structural use of concrete, Part 1, Code of practice for design and construction; Part
2, Code of practice for special circumstances. British standard institute, BSI, London, 1985.

51
Chapter 3: Hardened concrete properties

CHAPTER 3

HARDENED CONCRETE PROPERTIES


Chapter 3: Hardened concrete properties

Table of Contents

3.1. INTRODUCTION ........................................................................................................................ 52


3.2. MANUFACTURING PROCESS OF EXPANDED POLYSTYRENE BEADS .......................... 54
3.2.1. Pre-Expansion Stage............................................................................................................. 54
3.2.2. Aging Stage .......................................................................................................................... 54
3.2.3. Moulding Stage .................................................................................................................... 54
3.3. EXPERIMENTAL DATA FROM LITERATURE REVIEW ...................................................... 56
3.3.1. Mixture Design of Lightweight Concrete Containing Expanded Polystyrene Beads ........... 57
3.4. MECHANICAL PROPERTIES OF LIGHTWEIGHT CONCRETE CONTAINING
EXPANDED POLYSTYRENE BEADS ................................................................................................... 69
3.4.1. Modulus of Elasticity ........................................................................................................... 77
3.4.1.1. Experimental and analytical database for modulus of elasticity .................................. 77
3.4.1.2. Proposed model of modulus of elasticity in this study ................................................ 78
3.4.1.3. Comparison of the analytical models of modulus of elasticity .................................... 79
3.4.2. Modulus of Rupture in Lightweight Concrete ...................................................................... 83
3.4.3. Comparison of the Analytical Models of Modulus of Rupture ............................................ 84
3.4.4. Splitting Tensile Strength in Lightweight Concrete ............................................................. 86
3.4.5. Relationship between Modulus of Rupture and Modulus of Elasticity ................................ 89
3.4.6. Relationship between splitting tensile strength and modulus of elasticity ........................... 89
3.5. MECHANICAL PROPERTIES OF CONVENTIONAL AND SELF-COMPACTING
CONCRETE............................................................................................................................................... 90
3.5.1. Conventional Concrete ......................................................................................................... 90
3.5.1.1. Modulus of elasticity ................................................................................................... 90
3.5.1.2. Modulus of rupture ...................................................................................................... 94
3.5.1.3. Splitting tensile strength .............................................................................................. 94
3.5.2. Self-compacting Concrete .................................................................................................... 95
3.5.2.1. Modulus of elasticity ................................................................................................... 95
3.5.2.2. Modulus of rupture ...................................................................................................... 97
3.5.2.3. Splitting tensile strength .............................................................................................. 97
3.6. COMPRESSIVE STRESS-STRAIN CURVE IN LIGHTWEIGHT, CONVENTIONAL AND
SELF-COMPACTING CONCRETE ......................................................................................................... 98
3.6.1. Compressive Stress-Strain Curve in Lightweight Concrete ............................................... 101
3.6.2. Compressive Stress-Strain Models in Conventional Concrete ........................................... 103
3.6.3. Compressive Stress-Strain Models in Self-compacting Concrete ...................................... 105
3.7. STEEL-CONCRETE BOND CHARACHTERISTICS .............................................................. 107
3.7.1. Bond Characteristics in Lightweight Concrete ................................................................... 107
3.7.2. Bond Characteristics in Conventional Concrete ................................................................. 107
Chapter 3: Hardened concrete properties

3.7.3. Bond Characteristics in Self-compacting Concrete ............................................................ 109


3.8. SUMMARY ................................................................................................................................ 110
REFERENCES ......................................................................................................................................... 112
Chapter 3: Hardened concrete properties

List of Figures

Figure 3.1: Comparison of the experimental data with the predicted modulus of elasticity in EPS-LWC 79
Figure 3.2: Comparison of the existing MoE models with the proposed models (Eq.3.1) (density 2000
kg/m3) ......................................................................................................................................................... 82
Figure 3.3: Comparison of the existing MoE models with the proposed models (Eq. 3.1) (density 1280
kg/m3) ......................................................................................................................................................... 83
Figure 3.4: Comparison of the modulus of rupture in experimental data and prediction of Eq. (3.2) ........ 84
Figure 3.5: Comparison of the existing MoR models with the proposed models (density 1280 kg/m 3) .... 85
Figure 3.6: Comparison of the experimental and predicted splitting tensile strength in EPS-LWC .......... 87
Figure 3.7: Comparison of the proposed and existing models of splitting tensile strength (density 2000
kg/m3) ......................................................................................................................................................... 89
Figure 3.8: Different types of concrete failure in compression ................................................................. 99
Figure 3.9: Comparing the failure modes of LWC and CC [166] ............................................................ 100
Chapter 3: Hardened concrete properties

List of Tables

Table 3.1: Physical properties requirements of EPS, according to AS 1366, Part 3 – 1992 ...................... 55
Table 3.2: Other physical properties of EPS not specified in AS1366, Part 3-1992 .................................. 55
Table 3.3: Physical properties of Expanded Poly-Styrene (EPS) in ASTM C-578-15 .............................. 56
Table 3.4: Source of collected data for LWC containing EPS beads ......................................................... 58
Table 3.5: Mixture design of EPS light weight concrete from literature .................................................... 63
Table 3.6: Investigated characteristics of LWC containing EPS beads in literature .................................. 70
Table 3.7: Symbols and abbreviations for measured properties of EPS-LWC in Table 3.3 ...................... 72
Table 3.8: Mechanical properties of LWC containing EPS beads from experiments in literature ............. 73
Table 3.9: Coefficients of equation (3.1) and the compressive strength limit ............................................ 79
Table 3.10: Existing models to predict MoE in LWC ................................................................................ 80
Table 3.11: Coefficients of Equation (3.2) and the compressive strength limit ......................................... 84
Table 3.12: Existing models to predict the modulus of rupture of lightweight concrete ............................ 85
Table 3.13: Coefficients of equation (3.3) and the compressive strength limit .......................................... 86
Table 3.14: Existing models to predict the splitting tensile strength of lightweight concrete .................... 88
Table 3.15: Coefficients of Equation (3.4) and compressive strength limit ............................................... 89
Table 3.16: Coefficients of Equation (3.5) ................................................................................................. 90
Table 3.17: Existing models in codes of practice to predict MoE of conventional concrete ...................... 91
Table 3.18: Empirical models to predict the modulus of elasticity of conventional concrete .................... 93
Table 3.19: Existing models to predict the modulus of rupture of conventional concrete ......................... 94
Table 3.20: Existing models to predict the splitting tensile strength of conventional concrete ................. 95
Table 3.21: Modulus of elasticity of SCC- models in codes of practice and empirical equations ............. 96
Table 3.22: Modulus of rupture of SCC models in codes of practice and empirical equations ................. 97
Table 3.23: Splitting tensile strength in codes of practice and experimental studies in SCC ..................... 98
Table 3.24: Existing models of compressive stress-strain behaviour in LWC ......................................... 102
Table 3.25: Existing models of compressive stress-strain behaviour in CC ............................................ 103
Table 3.26: Existing models of compressive stress-strain curve for SCC ................................................ 106
Table 3.27: Models to estimate the steel-concrete bond in light-weight concrete .................................... 107
Table 3.28: models to estimate the steel-concrete bond in conventional concrete ................................... 108
Table 3.29: Models to estimate the steel-concrete bond in self-compacting concrete ............................. 110
Table 3.30: Proposed models for structural EPS lightweight concrete .................................................... 111
Chapter 3: Hardened concrete properties

Chapter 3: Hardened concrete properties

3.1. INTRODUCTION

Mechanical properties of concrete change from a fresh to hardened state with wholly
different trend and values. Various investigations have tried to find and establish a relationship
between these two phases. The principal properties of hardened concrete which are of practical
importance can be listed as following:
- Strength;
- Permeability and durability;
- Shrinkage and creep deformation;
- Response to temperature variation.

The compressive strength is the most important property of concrete and almost all other
properties are defined in terms of compressive strength. Concrete is usually used for
compressive loads, and the compressive strength is a reliable measure of other properties. The
compressive strength of concrete is the most common measure for judging not only the ability
of the concrete to withstand load, but also the quality of the hardened concrete.

Durability might be defined as the capacity to maintain satisfactory performance over the
extended service life. A durable concrete withstands to a satisfactory degree the effects of
service conditions to which it will be subjected. The durability of concrete can be defined also
as its resistance to deterioration resulting from external and internal causes. In general, external
conditions such as environment and internal conditions like permeability, mix proportions and
entrapped air affect the durability of concrete.

52
Chapter 3: Hardened concrete properties

Fully-cured hardened concrete must be strong enough to withstand the structural and service
loads which will be applied to it and must be durable enough to withstand the environmental
exposure for which it is designed. If concrete is made with high-quality materials and is
properly proportioned, mixed, handled, placed and finished, it will be the strongest and most
durable building material

In Chapter Two, a brief literature review of the mechanical and time-dependent properties of
Light-Weight Concrete (LWC) is presented. Also, a general comparison between the
characteristics of LWC and Conventional Concrete (CC) is performed.

The overall objective of this study is to compare the instantaneous and time-dependent
deflection of one-way slabs made with LWC, CC and self-compacting concrete (SCC).
Therefore, in this Chapter, a comparative study of the mechanical properties of LWC, SCC and
CC is performed. In this regard, the existing models of international codes of practice and the
empirical models developed by researchers to predict the properties of LWC, CC and SCC are
compared.

Water, aggregates and cement are the main components of the mixture design of CC, LWC
and SCC. However, due to considerable differences in the proportioning of these components,
along with the inclusion of different types and volumes of mineral and chemical additives,
sometimes there are significant differences in the final composition and the fresh properties of
each type of concrete. The major differences can also be seen in the hardened properties of
LWC, SCC and CC. Compared to CC and SCC, LWC shows less density and usually lower
values of the basic mechanical properties, while SCC is preferred to CC and LWC, in term of
performance and the workability requirements of fresh concrete.

Lightweight concrete as described in Chapter Two, involves different types of natural and
artificial lightweight aggregates. In this study, a specific type of lightweight aggregate called
Expanded Poly-Styrene (EPS) has been used in the LWC mixture. Therefore, a particular
review of the application, mixture design and mechanical properties of LWC made with EPS
beads is performed. The accurate estimation of essential mechanical properties of EPS-LWC,
especially for structural purposes is vital in its application. Hence, based on a massive range of
experimental data from literature and the current study, new models are developed to predict the
modulus of elasticity, modulus of rupture, splitting tensile strength and fracture energy, stress-
strain curve and the steel-concrete bond behaviour for this type of LWC.

The proposed models are compared to the corresponding existing models of LWC with other
types of lightweight aggregates. Also, they are compared to the relevant models for CC and
SCC. The time-dependent characteristics like creep and shrinkage strain are evaluated in
Chapter Four.
53
Chapter 3: Hardened concrete properties

3.2. MANUFACTURING PROCESS OF EXPANDED POLYSTYRENE


BEADS

Expanded polystyrene beads are manufactured in three main steps of pre-expansion, aging
and moulding. A brief summary of each step is given in the following section.

3.2.1. Pre-Expansion Stage

The small expandable beads are subjected to steam, which causes the thermoplastic
polystyrene to soften. Increasing vapour pressure caused by the blowing agent causes the beads
to expand to up to 40 times their original volume. The pre-foam stage determines the final
density of the expanded polystyrene block.

3.2.2. Aging Stage

After pre-expansion, the pre-foam is transferred via fluidized drying bed to large silos for
aging. This process is designed to allow for the replacement of the expanding agent by air in the
cells of the bead. Aging also provides stabilization and cooling of the pre-foam.

3.2.3. Moulding Stage

Once conditioned by aging, the pre-foam is blown into a mould where further steaming
causes the expanded beads to fuse into a block.

Table (3.1) shows the chemical and mechanical properties of expanded polystyrene according
to ASTM-C/578 (2015) [1]. Table (3.2) also gives some information about the physical
properties of EPS beads not specified in Australian standard, AS-1366.3(1992) [2].

Table (3.3) describes the physical properties of EPS in the American standard ASTM-C578-15.

54
Chapter 3: Hardened concrete properties

Table 3.1: Physical properties of expanded polystyrene according to AS 1366, Part 3 – 1992

Classes Test
Physical Property Unit
L SL S M H VH Method

Nominal Density (kg/m3) 11 13.5 16 19 24 28

Compressive stress at 10% deformation (min) KPa 50 70 85 105 135 165 AS-2498.3

Cross-breaking strength (min) KPa 95 135 165 200 260 320 AS-2498.4

Rate of water vapor transmission (max)


μg/m2s 710 630 580 520 460 400 AS-2498.5
measured parallel to rise at 23°C

Dimensional stability of length, width, thickness


% 1.0 1.0 1.0 1.0 1.0 1.0 AS-2498.6
(max) at 70°C, dry condition 7 days

AS-2464.5
Thermal resistance (min) at a mean temperature or
M2K/W 1 1.13 1.17 1.2 1.25 1.28
of 25°C (50mm sample)
AS-2464.6

Flame propagation characteristics:


- median flame duration; max S 2 2 2 2 2 2
- eighth value; max S 3 3 3 3 3 3 AS-2122.1
- median volume retained; % 15 18 22 30 40 50
- eighth value (min.) % 12 15 19 27 37 47

Table 3.2: Other physical properties of expanded polystyrene not specified in AS1366, Part 3-1992

Properties Class
Test
Physical Property Unit L SL S M H VH
method
Density - Nominal Kg/m3 11 13.5 16 19 24 28
Compressive stress
at 1% deformation kPa 14 23 31 42 58 72 ASTM
(max.)

Compressive stress
at 5% deformation kPa 33 59 68 95 134 164 ASTM
(max.)

Flexural strength ASTM


kPa 60 150 178 218 304 337
(min.) C203

Elastic modulus
kPa 1450 2200 3100 4250 5850 7250
(min.)

Water absorption by ASTM


total immersion Volume% 4.0 4.0 4.0 3.0 30 2.0
(max.) C272

Buoyancy force Kg/m3 989 986.5 986.5 981 976 972

Coefficient of ASTM
mm/mm 0K 6.3e-6 6.3e-5 6.3e-6 6.3e-5 6.3e-6 6.3e-5
thermal expansion D696

55
Chapter 3: Hardened concrete properties

Table 3.3: Physical properties of expanded polystyrene in ASTM C-578-15

Type of EPS (classified by ASTM C-578)


ASTM
Property Type Type Type Type Type
test
XI I VIII II IX

Density, kg/m3
C303/
-Nominal 12 16 20 24 32
D1622
Minimum 11 15 18 22 29

Compressive strength, KPa


-at 0.5% strain C165/ 17 24 29 41 55
-at 1.0% strain D1621 35 48 58 82 110
-at yield point or 10% strain 355 69 90 104 173

Flexural strength, KPa C203 70 173 208 276 345

C272
Water absorption, Maximum % by
(Total 4.0 4.0 3.0 3.0 2.0
volume
immersion)

Modulus of elasticity, KPa D1621 3103 4655 5862 7935 10344

Poison’s ratio 0.05 0.05 0.05 0.05 0.05

3.3. EXPERIMENTAL DATA FROM LITERATURE REVIEW

The database of experimental results from published investigations could be one of the
effective tools to propose and verify new models and to compare the real and predicted values.
Accurate application of a developed model to particular concrete mixtures needs to use the
experiment data consistent with the applied testing methodology. In spite of effectiveness of
experimental results from different sources, using them can be problematic owing to the
following parameters:

a) Insufficient information concerning the exact composition of the concrete mixes;

b) Different size and number of the specimens, curing conditions and testing methodology;

c) Extracting real data of experimental results from graphs and diagrams.

The collected experimental database of this study has been based on papers presented at
conferences and published articles on LWC containing EPS beads. The database contains
information about the composition of the mixes, type of chemical admixture such as plasticizer
and air entraining agents, curing method, curing age, fine and coarse aggregate, filler type,
cement type, and fresh and hardened properties of LWC like density and compressive strength
at different ages.

56
Chapter 3: Hardened concrete properties

The wide range of experimental data of different mechanical properties of LWC containing
EPS beads is included in this study. These data are collected from 55 published papers in
journals and conference proceedings, some professional and scientific reports and also the
dissertations all around the world since 1976 to 2015. Table (3.4) shows the sources of data and
some general information about the collected data. The number of mixtures for LWC containing
EPS beads, curing methods, type and maximum size of the natural and lightweight course and
fine aggregates, maximum size of EPS beads, additional chemical admixtures including the
Super Plasticiser (SP), High Range Water Reducing Admixture (HRWRA) and Air-Entraining
Agent (AEA) are illustrated in the table. In addition, the types of fibre used in some
experimental studies in the literature are given.

3.3.1. Mixture Design of Lightweight Concrete Containing Expanded


Polystyrene Beads

In the LWC containing EPS beads, the coarse aggregates are replaced partially or wholly by
EPS beads. Table (3.5) presents the mixture components and proportions in some experimental
investigations to reach LWC. The resultant fresh density and compressive strength of the
mixture at age 28 days are also presented in the table.

57
Chapter 3: Hardened concrete properties

Table 3.4: Source of collected data for LWC containing EPS beads
Aggregates
Cementitious
EPS Chemical admixtures
No. of materials Coarse Fine
Curing bead Mineral
References EPS Fibre
method Max Max Max Filler
mixes
Cement Other Type Size Type size size SP HRWRA AEA
(mm) (mm) (mm)
MB-VR
neutralize
Bagon and Yannas W, Ottawa 2.36, Pozzolith
11 PC-II Perlite d Vinsol
(1976) [3] SSS sand 4.75 200-N
Resin
solution
Cormix
Bischoff el al. (1990) [4] 3 DH OPC sand 1.2 2.5-5.5
SP5
Ravindrarajah and Tuck crushed
6 W, CS, CR GP-PC SF 10 Sand 2.36 SF N.G. N.G.
(1994)[5] basalt
Park and Chisholm Puketapu
6 LSW GP-PC 2.75 N.G. FA
(1999) [6] sand
crushed
Babu and Babu (2003) SW, WGB, 1.18, 4.75,
11 PC-I SF blue 10, 16 sand SF NB
[7] W 2.36 6.3
granite
crushed
Babu and Babu (2004) Immersion FA, 4.75,
7 PC-I blue 8 sand 2.36 SF NB N.G.
[8] (SW, W) SF 6.3
granite
crushed River
Chen and Liu (2004) [9] 12 FR OPC SF 20 2.85 3, 8 SF NB SF
granite sand
Oven(60 Saint Betocarb Optima Meyco-
Miled et al. (2004) [10] 2 o N.G. 3,5
C), CR Vigor P2 100 685
Chryso
CEM I SF rounded 400μ
Roy et al. (2005) [11] 5 Oven(60 oC) 1, 3, 6 SF THP
52.5 (A) quartz m
(PCB)
crushed
FA 4.75,
Babu et al. (2005) [12] 7 WGB, W PC-I-C53 blue 8 Sand 2.36 NB N.G.
(F) 6.3
granite
styrene-
crushed butadiene
Chen and Liu (2005) [13] 24 NAD, W, M OPC 20 River sand N.G. 3, 8 N.G.
granite rubber
(SBR)
(crushed
FA(F), 1.18,
Babu et al. (2006) [14] 6 N.G. OPC blue 8 Sand 2.36 N.G.
SF 4.75
granite

58
Chapter 3: Hardened concrete properties

Table 3.4: Source of collected data for LWC containing EPS beads (continued)

Aggregates
Cementitious
EPS Chemical admixtures
No. of materials Coarse Fine
Curing bead Mineral
References EPS Fibre
method Max Max Max Filler
mixes
Cement Other Type Size Type size size SP HRWRA AEA
(mm) (mm) (mm)
PA-
Haghi et al. (2006) [15] 41 NAD, W OPC N.G. N.G. 3, 6, 8 NB
66F
Bouvard et al. (2007) [16] 8 N.G. HPC N.G. N.G. N.G. N.G. 1, 2.2 N.G.
styrene-
crushed
River butadiene
Chen and Liu (2007) [17] 24 NAD, W, M OPC granite 20 N.G. 3, 8 N.G.
sand rubber
(SBR)
NAD,
Wrapped in CEM I rounded 300μ 1, 2.5, Optima
Miled et al. (2007) [18] 5 SF(A)
aluminium 52.5 quartz m 6.3 175 (PCB)
paper (CR)
Amianti and Botaro humid
4 HSC gravel 6.88 Sand 4.8 N.G.
(2008) [19] chamber
0.5, 1,
Bisschop and Mier (2008) CEM I glass quartz,
33 N.G. 2 N.G. N.G. N.G.
[20] 52.5R beads feldspar
, 4, 6
300μ ASTM
Lepech et L. (2008) [21] 3 N.G. OPC FA(F) Silica N.G. PVAF
m C1017
Chen and Fang (2009) PPF,
12 SC, FR PC-I SF Sand 2.36 600μm SF NB
[22] SF
SF, FDN
Bai et al. (2011) [23] 5 N.G. OPC stone N.G. sand N.G. 2-8
FA (NB)
Chen and Fang (2011) 1, 2.5,
3 FR PC-I SF Sand 2.36 NB PPF
[24] 6.3
River GLENIU
Ling and Teo (2011) [25] 4 NAD, W PC-I RHA 1.18 8 RHA
sand M C380
Polycarbox
SF,
Madandoust et al. (2011) crushed River ylic ether
12 N.G. PC-I Nano- 12.5 4.75 5,8 SF
[26] gravel sand based(Glen
SiO2
ium51)
NAD,
Wrapped in CEM I rounded 300μ 1, 2.5, Optima
Miled et al. (2011) [27] 5 SF (A)
aluminium 52.5 quartz m 6.3 175 (PCB)
paper (CR)

59
Chapter 3: Hardened concrete properties

Table 3.4: Source of collected data for LWC containing EPS beads (continued)

Cementitious
Aggregates
materials
Chemical admixtures
No. of EPS
Curing Coarse Fine Mineral
References EPS bead Fibre
method Filler
mixes Cement Other Max Max Max
Type Size Type Size Size SP HRWRA AEA
(mm) (mm) (mm)
Sadrmomtazi et al. SF,
12 Wet cured PC-I gravel 12 Sand 4.75 3.5 PCB PPF
(2012) [28] RHA
crushed River
Xu et al. (2012) [29] 9 CR OPC 10 N.G. 3
stone sand
Carbonari et al. (2012) CEM II- Moravac
1 W 4 5
[30] 32.5R sand
Ferrandiz-Mas and Rheomix Hidroxipro Basf
CEM I Silica
Garcia-Alcocel (2012) 12 W N.G. 1,3 GT 205 pil metil Rheomix
52,5R sand
[31] MA cellulose a 934,
Liu and Guo (2012) [32] 7 W PC-I SF 5 Sand N.G. N.G.
white
Passa et al. (2012) [33] 3 W HSP 4 N.G.
marble
FDN
Zhao et al. (2012) [34] 5 N.G. N.G. SF stone Sand N.G. 2-8 SF
(NB)
crushed
Trussoni et al. (2013) plastic
2 PC-I limeston 10, 19 sand N.G. 1
[35] cover, M
e
hydroph Coagulant, Silane
Wang and Fu (2013) FA, Gum hydrophob
7 N.G. RHSC obic N.G. 2.5,5 HPMC coupling PPF
[36] MCF powder ic agent
perlite ether agent
F-7000 animal
Chen and Liu (2013) PC-I,
4 WGB, FR SF 1 SF NB latex protein- PPF
[37] HAC
powder based
crushed
Elsalah et al. (2013) [38] 6 Oven, W OPC 5 Sand N.G. N.G.
stone
Basf
Ferrandiz-Mas and Basf
CEM I silica Rheomix
Garcia-Alcocel (2013) 15 W 1,3 HPMC Rheomix
52.5R sand GT 205
[39] 934
MA
plastic Natural
Herki et al. (2013) [40] 6 OPC FA 3 10 FA
cover, W sand
Sadrmomtazi et al. (2013) SF, crushed
9 W OPC 12 River sand 4.75 3.5 PCB PPF
[41] RHA limestone

60
Chapter 3: Hardened concrete properties

Table 3.4: Source of collected data for LWC containing EPS beads (continued)
Aggregates
Cementitious
EPS Chemical admixtures
No. of materials Coarse Fine
Curing bead Mineral
References EPS Fibre
method Max Max Max Filler
mixes
Cement Other Type Size Type size size SP HRWRA AEA
(mm) (mm) (mm)
Hongbo et al. (2014) Standard
7 N.G. N.G. 1-5
[42] curing
Hidayat and Gunaedi
18 N.G. PC-I FA Sand N.G. 1-2
(2014) [43]
Kekanović et al. (2014) CEM II- Moravac
1 W 4 5 PPF
[44] 32.5R sand
Liu and Chen (2014) 600
8 SC, FR PC-I SF Sand 2.36 NB
[45] μm, 4
Mousavi et al. (2014) Sand
2 NAD Fly ash 5 GF
[46] powder
Schackow et al. (2014) PC-II- Vermic River Vermicul Glenium-
8 N.G. 3 4.75 3
[47] Z32 ulite sand ite SCC160
crushed River
Tamut et al. (2014) [48] 6 W OPC N.G. N.G. 8-9
granite sand
SF,
Nekooie et al. (2015) crushed
4 N.G. OPC FA, 12 Sand N.G. 2-5
[49] stone
POFA
MEDI
crushed siliceous UM
Xu et al. (2015) [50] 16 N.G. OPC 10 3
granite sand GRA
DED
>2.36,
>1.18,
Li et al. (2015) [51] 3 CB MPC FA N.G. >4.75 >1.18
>600μ
m
Lo Monte et al. (2015) CEM II
2 CR gravel 12 Sand N.G. 3 N.G. N.G.
[52] R 42.5
acrylic-
based Viscofluid Mapeplast
CEM II Natural
Pecce et al. (2015) [53] 2 M SF gravel 20 N.G. 3 (Dynamon SCC/10, PT1,
R 42.5 sand
SX, Mapei Mapei
Mapei)
Ranjbar and Mousavi crushed River 4.75, Glenium5
W PC-I SF 12.5 4.75
(2015) [54] gravel sand 9.5 1

61
Chapter 3: Hardened concrete properties

The notations and symbols used in Table (3.4) are illustrated below.

Curing method

Water (W), Fog Room (FR), Natural Dry (ND), Sodium Sulphate Solutions (SSS), Damp
Hessian (DH), Chemical Solutions (CS), Curing Room (CR), Moist (M), Curing Box (CB),
Natural Air Dried (NAD), Wet Gunny Bags (WGB), Sea Water (SW), Stripped Cover (SC),
Lime Saturated Water (LSW)

Cementitious materials and mineral fillers

Cement: Ordinary Portland Cement (OPC), Portland Cement type 1 (PC-I), type 2 (PC-II),
Portland Cement type 2 grade 42.5 (CEM-II-R 42.5), Portland Cement type 1 class 53 (PC-I
C53), General Purpose (GP), High Performance Cement (HPC), High Strength Cement (HSC),
High Alumina Cement (HAC), Magnesia Phosphate Cement (MPC), Rapid Hardening
Sulphoaluminate Cement (RHSC), High-Performance cement (HPC)

Other:

Silica Fume (SF), Fly Ash (FA), Rice Husk Ash (RHA), Palm Oil Fuel Ash (POFA), Micro-
Cellular Foam (MCF)

Aggregates: Expanded Poly-Styrene (EPS)

Chemical admixtures:

Super Plasticiser (SP): Naphthalene-Based (NB), Poly-Carboxylate-Based (PCB)

High Range Water Reducing Admixture (HRWRA):

Air-Entraining Agent (AEA): Hydroxy-Propyl Methyl Cellulose (HPMC)

EPS: Expanded Poly-Styrene aggregate (EPS bead)

Fibre type: Glass Fibre (GF), Poly-Propylene Fibre (PPF), Steel Fibre (SF), Poly-Vinyl
Alcohol Fibre (PVAF), Poly-Amide-66 Fibre (PA-66F)

N.G. in all columns and rows mentioning that the method or material is applied, but the relevant
information is not given
*
In Sadrmomtazi et al. (2012) [28], aggregates 0-3 mm are classified as sand and aggregates in
the range of 3-6 mm and 6-12 mm are considered as coarse aggregates respectively.

62
Chapter 3: Hardened concrete properties

Table 3.5: Mixture design of EPS lightweight concrete from literature

Reference cement Silica RHA Fly Nano- Coarse sand EPS Water w/c SP AEA Fibre γ ࢌᇱࢉ
fume ash sio2 aggregate MPa
3 3 3 3 3 3 3 3 3 3
kg/m kg/m kg/m kg/m kg/m kg/m kg/m Lit/m kg/m ml/kg ml/kg Vol.% kg/m MPa
Ravindrarajah 400 0 0 753 690 l/m3 140 0.35 1325 11.9
and Tuck 400 0 0 713 775 160 0.4 1230 9.8
(1994)[5] 400 0 0 673 710 200 0.5 1220 8.7
400 0 0 633 775 240 0.6 1165 5.6
400 40 250 507 675 176 0.4 1305 11.8
360 40 250 563 675 160 0.4 1350 10.9
Park and 350 0 0 1000 l/m3 180 0.51 550 0.8
Chisholm 230 120 0 1000 180 0.51 520 0.7
(1999) [6] 400 0 180 900 180 0.45 820 3.8
270 130 180 900 180 0.45 760 2.1
450 0 340 850 180 0.4 1020 6.7
300 150 340 850 180 0.4 980 4.9
Babu and 350 3% 325 481 7.07(36.4%) 158 0.44 0.25% 1552 10.2
Babu (2003) 362 3% 481 488 5.63(29%) 164 0.44 0.25% 1698 15
[7] 375 3% 648 497 4.07(21%) 170 0.44 0.25% 1979 14
335 3% 779 6.72(34.4%) 152 0.44 0.25% 1503 8.4
399 5% 530 498 7.2(22.2%) 185 0.44 0.25% 1856 15.2
399 5% 1028 7.2(22.2%) 185 0.44 0.25% 1748 17.6
399 5% 743 10.8(33.2%) 185 0.44 0.25% 1546 12.4
399 5% 515 337 7.3(28.7%) 185 0.44 0.25% 1762 11.6
399 5% 515 337 7.3(28.7%) 185 0.44 0.25% 1771 14.2
605 9% 466 305 6.6(26.3%) 185 0.278 0.25% 1793 15.6
605 9% 466 305 6.6(26%) 185 0.278 0.25% 1873 19.8

63
Chapter 3: Hardened concrete properties

Table 3.5: Mixture design of EPS lightweight concrete from literature (continued)

Silica Fly Nano- Coarse ࢌᇱࢉ


cement RHA sand EPS Water SP AEA Fibre γ
Reference fume ash sio2 aggregate w/c MPa
3 3 3 3 3 3 3 3 3 3
kg/m kg/m kg/m kg/m kg/m kg/m kg/m Lit/m kg/m ml/kg ml/kg Vol.% kg/m MPa
3
472 0 710 392 2.49kg/m (25%) 175 0.37 4.2 0 1820 22.1
472 0 710 392 2.49 (25%) 175 0.37 4.2 70 1883 21.4
472 0 455 255 5.01 (40%) 175 0.37 4.5 0 1356 17.6
472 0 455 255 5.01 (40%) 175 0.37 4.5 70 1403 16.7
472 0 201 118 6.88 (55%) 175 0.37 5.1 0 876 10.6
Chen and Liu 472 0 201 118 6.88 (55%) 175 0.37 5.1 70 910 9.9
(2004) [9] 472 35 710 392 2.49 (25%) 175 0.37 4.2 0 1850 25.7
472 35 710 392 2.49 (25%) 175 0.37 4.2 70 1929 25.9
472 35 455 255 5.01 (40%) 175 0.37 4.5 0 1370 20.1
472 35 455 255 5.01 (40%) 175 0.37 4.5 70 1415 20.8
472 35 201 118 6.88 (55%) 175 0.37 5.1 0 882 11.3
472 35 201 118 6.88 (55%) 175 0.37 5.1 70 392 10.9
287 86 229 23.31kg/m3 74.5 0.26 1.3% 600
330 99 264 21.86 85.7 0.26 1.3% 800
Roy et al.
416 125 333 19.05 108.1 0.26 1.3% 1000
(2005) [11]
502 150 401 16.05 130.4 0.26 1.3% 1200
588 176 470 13.14 152.8 0.26 1.3% 1400
142 142 5.5% 66.5% 185 0.651 582 1.1
190 190 10% 58% 185 0.487 779 2.3
Babu et al. 224 224 20% 49% 185 0.431 984 3.83
(2005) [12] 247 247 16% 20% 38% 214 0.434 1304 6
274 274 30% 20% 28.5% 225 0.412 1484 7.8
309 309 50% 20% 16.3% 245 0.396 1723 12.5

64
Chapter 3: Hardened concrete properties

Table 3.5: Mixture design of EPS lightweight concrete from literature (continued)

Silica Fly Nano- Coarse ࢌᇱࢉ


Reference cement RHA sand EPS Water SP AEA Fibre γ
fume ash sio2 aggregate w/c MPa
3 3 3 3 3 3 3 3 3 3
kg/m kg/m kg/m kg/m kg/m kg/m kg/m Lit/m kg/m ml/kg ml/kg Vol.% kg/m MPa
394 169 15% 50% 200 0.355 0.5% 1012 7
394 169 26% 14% 35% 200 0.355 0.5% 1492 11
Babu et al. 394 169 42% 24% 20% 200 0.355 0.5% 1858 20
(2006) [14] 394 169 15% 50% 200 0.355 0.5% 1076 5
394 169 26% 14% 35% 200 0.355 0.5% 1454 9
394 169 42% 24% 20% 200 0.355 0.5% 1842 17
7.24
839.35 251.83 685.86 2.55(12.8%) 231.17 2000
kg/m3
754 226.39 616.58 4.31(21.7%) 191.64 6.51 1800
669.78 200.96 547.3 6.07(30.5%) 170.11 5.78 1600
585 175.52 478.02 7.84(39.4%) 148.58 5.05 1400
500.22 150.08 408.74 9.6(48.3%) 127.04 4.32 1200
839.35 251.83 685.86 2.55(12.8%) 231.17 7.24 2000
Miled et al. 754 226.39 616.58 4.31(21.7%) 191.64 6.51 1800
(2007) [18] 669.78 200.96 547.3 6.07(30.5%) 170.11 5.78 1600
585 175.52 478.02 7.84(39.4%) 148.58 5.05 1400
500.22 150.08 408.74 9.6(48.3%) 127.04 4.32 1200
839.35 251.83 685.86 2.55(12.8%) 231.17 7.24 2000
754 226.39 616.58 4.31(21.7%) 191.64 6.51 1800
669.78 200.96 547.3 6.07(30.5%) 170.11 5.78 1600
585 175.52 478.02 7.84(39.4%) 148.58 5.05 1400
500.22 150.08 408.74 9.6(48.3%) 127.04 4.32 1200

65
Chapter 3: Hardened concrete properties

Table 3.5: Mixture design of EPS light weight concrete from literature (continued)

Silica Fly Nano- Coarse ࢌᇱࢉ


cement RHA sand EPS Water SP AEA Fibre γ
Reference fume ash sio2 aggregate w/c MPa
kg/m3 kg/m3 kg/m3 kg/m3 kg/m3 kg/m3 kg/m3 Lit/m3 kg/m3 ml/kg ml/kg Vol.% kg/m3 MPa
437 77 321 18.1(55%) 144 0.28 0.8 1% 980 8.5
728 128 535 4.95(15%) 240 0.28 1.1 1% 1630 44
631 111 464 11.5(35%) 208 0.28 1.1 1% 1435 23.5
437 77 321 18.1(55%) 144 0.28 1 1% 990 9.5
Chen and Fang 728 128 535 4.95(15%) 240 0.28 1.2 1% 1590 46
(2009) [22] 631 111 464 11.5(35%) 208 0.28 1.2 1% 1418 24.5
437 77 321 18.1(55%) 144 0.28 1.2 1% 975 10
728 128 535 4.95(15%) 240 0.28 1 1% 1610 46.8
631 111 464 11.5(35%) 208 0.28 1 1% 1430 25
437 77 321 18.1(55%) 144 0.28 1 1% 983 10.5
386 29.68 213.5 899 503 10% 184 5.93
386 29.68 213.5 728 407 20% 184 5.93
Bai et al.
386 29.68 213.5 557 311 30% 184 5.93
(2011) [23]
386 29.68 213.5 386 215 40% 184 5.93
386 29.68 213.5 215 119 50% 184 5.93
861 369 768 246 0.2 1.2 1800 130
944 236 737 295 0.25 1 1400 90
1017 113 706 339 0.3 0.8 1000 50
861 369 246 1.2 1800
Chen and Fang
944 236 737 295 1 1400
(2011) [24]
1017 113 706 339 0.8 1000
861 369 246 1.2 1800
944 236 737 295 1 1400
1017 113 706 339 0.8 1000

66
Chapter 3: Hardened concrete properties

Table 3.5: Mixture design of EPS lightweight concrete from literature (continued)

Silica Fly Nano- Coarse ࢌᇱࢉ


cement RHA sand EPS Water SP AEA Fibre γ
RefereRence fume ash sio2 aggregate w/c MPa
3 3 3 3 3 3 3 3 3 3
kg/m kg/m kg/m kg/m kg/m kg/m kg/m Lit/m kg/m ml/kg ml/kg Vol.% kg/m MPa
425 0 850 18 0.5 8 1830 13
403 21 850 18 0.5 8 1810 15
Ling and Teo
382 42 850 18 0.5 8 1790 17
(2011) [25]
360 64 850 18 0.5 8 1770 13
340 85 850 18 0.5 8 1750 11
400 44 3.5 662 783 1.46 196 2185
400 44 3.2 594 716 2 196 2040
400 44 2.8 491 617 2.52 196 1896
400 44 2.5 387 517 2.74 196 1712
388 44 3.5 662 783 1.46 196 2187
Madandoust et 388 44 3.2 594 716 2 196 2105
al. (2011) [26] 388 44 2.8 491 617 2.52 196 1898
388 44 2.5 387 517 2.74 196 1701
434 48 3.5 662 783 1.46 183 2235
434 48 3.2 594 716 2 183 2077
434 48 2.8 491 617 2.52 183 1919
434 48 2.5 387 517 2.74 183 1748
400 0 0 872 540 2.14 170 0.43 N.G. 1.32 1900 33
400 0 0 696 431 2.85 165 0.41 N.G. 4 1700 16.7
400 0 0 475 294 3.11 160 0.4 N.G. 6.6 1350 9.8
400 0 0 249 154 2.24 160 0.42 N.G. 13.26 979 5.4
360 40 0 848 524 2.08 175 0.44 N.G. 1.32 1900 35
Sadrmomtazi 360 40 0 682 422 2.79 175 0.44 N.G. 4 1650 24.4
et al. (2012) 360 40 0 456 282 3 170 0.43 N.G. 6.6 1300 10.2
[28] * 360 40 0 227 140 2.04 167 0.42 N.G. 13.26 949 2.9
320 0 80 750 470 1.85 205 0.51 N.G. 1.32 1850 22.4
320 0 80 623 385 2.54 205 0.51 N.G. 4 1600 10.6
320 0 80 395 245 2.59 200 0.5 N.G. 6.6 1250 7.1
320 0 80 133 82 55%1.19 210 0.53 N.G. 13.26 890 3.5
0.0231 0 0.0338 0.0411 0.000287 0.0113 0.49 0 0 1920 23.24

67
Chapter 3: Hardened concrete properties

Table 3.5: Mixture design of EPS lightweight concrete from literature (continued)

Silica Fly Nano- Coarse ࢌᇱࢉ


cement RHA sand EPS Water SP AEA Fibre γ
Reference fume ash sio2 aggregate w/c MPa
3 3 3 3 3 3 3 3 3 3
kg/m kg/m kg/m kg/m kg/m kg/m kg/m Lit/m kg/m ml/kg ml/kg Vol.% kg/m MPa
400 35% 15% 0.45 2060 20.77
400 38% 20% 0.5 1890 14.79
400 41% 25% 0.55 1720 11.22
Xu et al. (2012) 450 41% 15% 0.45 1880 17.86
450 35% 20% 0.5 1800 7.85
[29]
450 38% 25% 0.55 1930 15.68
500 38% 15% 0.45 1730 13.52
500 41% 20% 0.5 1950 18.56
500 35% 25% 0.55 1790 12.98
Trussoni et 0.0233 0 0.032 0.0411 0.000299 0.011 0.48 0 0 1920 24.99
al. (2013)
0.0231 0 0.0338 0.0411 0.000287 0.0113 0.49 0 0 1920 23.24
[35]
Lo Monte et 643 591 475 5.09 200 7.05 2.12 24.7
al. (2015)
815 310 620 5.32 220 9.11 2.73 26.3
[52]
Pecce et al. 815 367 734 5.32 220 0.27 9.11 2.73 31
(2015) [53] 843 591 475 5.09 193 0.3 7.05 2.12 20.5
400 44 662 783 10% 196 0.44 3.5 2185 27
400 44 594 716 15% 196 0.44 3.2 2040 22
Ranjbar and 400 44 491 617 22.5% 196 0.44 2.8 1896 19
Mousavi 400 44 387 517 30% 196 0.44 2.5 1712 17
434 48 662 783 10% 183 0.38 3.5 2235 31
(2015) [54]
434 48 594 716 15% 183 0.38 3.2 2077 26
434 48 491 617 22.5% 183 0.38 2.8 1919 22
434 48 387 517 30% 183 0.38 2.5 1748 19

68
Chapter 3: Hardened concrete properties

3.4. MECHANICAL PROPERTIES OF LIGHTWEIGHT CONCRETE


CONTAINING EXPANDED POLYSTYRENE BEADS

Mechanical properties of LWC containing EPS beads are evaluated in various experimental
investigations. Table (3.6) explains the investigated properties of EPS-LWC in the collected
publications in this study. Table (3.7) also describes the abbreviations of the characteristics and
measured properties of EPS-LWC by different testing methods.

69
Chapter 3: Hardened concrete properties

Table 3.6: Investigated characteristics of lightweight concrete containing EPS beads in literature

Column 1 2 2 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20
Properties ࢌᇱࢉ ࡱࢉ ࢌ࢘ ࢌࢉ࢚ σ-ε γ WRK ࣘࢉࢉ ࢿ࢙ࢎ WA (%) DUR EAF BS UPV TC SE EAP MM ࡾࢎ࢓ ࡾࢋ࢒ࢋࢉ
Bagon and Yannas (1976)
3 3 3 3 3 3
[3]
Bischoff el al. (1990) [4] 3 3 3
Ravindrarajah and Tuck
3 3 3 3 3 3 3 3
(1994)[5]
Park and Chisholm (1999)
3 3 3 3 3
[6]
Babu and Babu (2003) [7] 3 3 3 3 3 3
Babu and Babu (2004) [8] 3 3 3 3
Chen and Liu (2004) [9] 3 3 3 3 3
Miled et al. (2004) [10] 3 3
Roy et al. (2005) [11] 3 3 3 3 3 3 3
Babu et al. (2005) [12] 3 3 3 3 3 3 3
Chen and Liu (2005) [13] 3 3 3
Babu et al. (2006) [14] 3 3 3 3 3 3
Haghi et al. (2006) [15] 3 3 3
Bouvard et al. (2007) [16] 3 3 3 3
Chen and Liu (2007) [17] 3 3 3
Miled et al. (2007) [18] 3 3 3 3
Amianti and Botaro (2008)
3 3
[19]
Bisschop and Mier (2008)
3 3 3
[20]
Lepech et L. (2008) [21] 3 3
Chen and Fang (2009) [22] 3 3 3 3 3 3
Bai et al. (2011) [23] 3 3 3
Chen and Fang (2011) [24] 3 3 3 3
Ling and Teo (2011) [25] 3 3 3 3 3 3
Madandoust et al. (2011)
3 3 3
[26]
Miled et al. (2011) [27] 3 3
Sadrmomtazi et al. (2012)
3 3 3 3 3 3 3 3 3
[28]

70
Chapter 3: Hardened concrete properties

Table 3.6: Investigated characteristics of lightweight concrete containing EPS beads in literature (continued)

Column 1 2 2 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20
Properties ࢌᇱࢉ ࡱࢉ ࢌ࢘ ࢌࢉ࢚ σ-ε γ WRK ࣘࢉࢉ ࢿ࢙ࢎ WA (%) DUR EAF BS UPV TC SE EAP MM ࡾࢎ࢓ ࡾࢋ࢒ࢋࢉ
Xu et al. (2012) [29] 3 3 3 3
Ferrandiz-Mas and
3 3 3 3 3
Garcia-Alcocel (2012) [31]
Liu and Guo (2012) [32] 3 3 3 3 3
Passa et al. (2012) [33] 3 3 3
Zhao et al. (2012) [34] 3 3
Trussoni et al. (2013) [35] 3 3 3 3
Wang and Fu (2013) [36] 3 3 3 3 3
Chen and Liu (2013) [37] 3 3 3 3 3 3
Elsalah et al. (2013) [38] 3 3 3
Ferrandiz-Mas and
3 3
Garcia-Alcocel (2013) [39]
Herki et al. (2013) [40] 3 3 3
Sadrmomtazi et al. (2013)
3
[41]
Hongbo et al. (2014) [42] 3 3 3 3 3 3
Wang and Arellano (2014)
3 3 3 3
[55]
Gao et al. (2014) [56] 3 3 3 3
Hidayat and Gunaedi (2014)
3 3 3
[43]
Kekanović et al. (2014)
3 3 3 3 3 3 3
[44]
Liu and Chen (2014) [45] 3 3 3 3 3 3
Mousavi et al. (2014) [46] 3 3 3
Schackow et al. (2014) [47] 3 3 3 3 3
Tamut et al. (2014) [48] 3 3 3 3
Nekooie et al. (2015) [49] 3 3 3
Xu et al. (2015) [50] 3 3 3 3
Li et al. (2015) [51] 3 3 3 3 3 3
Lo Monte et al. (2015) [52] 3 3 3 3 3 3 3
Pecce et al. (2015) [53] 3 3 3 3 3
Ranjbar and Mousavi
3 3 3 3 3 3 3
(2015) [54]

71
Chapter 3: Hardened concrete properties

Full description and symbols for properties illustrated in the columns of Table (3.6) are
presented in Table (3.7).

Table 3.7: Symbols and notations for collected properties of EPS-LWC in Table 3.6

Column Measured property Abbreviation


1 Compressive strength ݂௖ᇱ
2 Modulus of elasticity ‫ܧ‬௖
3 Flexural tensile strength (modulus of rupture) ݂௥
4 Splitting tensile strength ݂௖௧
5 Stress-strain ߪെߝ
6 Density ߛ
7 Slump or workability WRK
8 Creep coefficient ߶௖௖
9 Shrinkage strain ߝ௦௛
10 Water absorption WA (%)
11 Durability test DUR
12 Energy absorption or fracture energy EAF
13 Bond-slip between concrete and bars BS
14 Ultrasonic Pulse Velocity UPV
15 Thermal conductivity or fire resistance TC
16 Size effect of specimen or aggregates SE
17 Entrapped air or porosity EAP
18 Moisture migration MM
19 Rebound value (Schmidt hammer) ܴ௛௠
20 Electrical resistivity ܴ௘௟௘௖

The experimental data from previous experiments in the literature is always a reliable method
of developing new models and evaluating the existing models to predict the mechanical
properties of concrete. The main difference between conventional concrete and LWC containing
EPS beads is replacement of some or the whole part of the coarse aggregates with various sizes
of EPS beads.
Table (3.8) presents the recorded values of mechanical properties of EPS-LWC in the hardened
state from previously conducted experiments presented in Table (3.4). Compressive strength
(݂௖ᇱ ), modulus of elasticity (‫ܧ‬௖ ), modulus of rupture (݂௥ ) and splitting tensile strength (݂௖௧ ) are
the main mechanical properties given in Table (3.8).

Table (3.8) also explains the volume of EPS beads in the LWC mixtures. The maximum size
of EPS beads and fine and coarse aggregates are previously illustrated in Table (3.4).

72
Chapter 3: Hardened concrete properties

Table 3.8: Mechanical properties of LWC containing EPS beads from experiments in literature

EPS ࢌᇱࢉ ࡱࢉ ࢌ࢘ ࢌࢉ࢚


References
Vol.% (MPa) (GPa) (MPa) (MPa)
3.3 3.92 1.37
4.6 3.92 1.76
4.6 3.92 1.66
4.6 4.89 1.96
3.9 3.92 1.57
Bagon and Yannas (1976) [3] 4.6 3.92 1.76
4.8 3.92 1.76
4.8 5.87 1.76
3.6 3.92 1.37
5 5.87 1.76
4.5 4.89 1.66
36.3 11.9 11.8 1.7
36.3 9.8 8.7 1.5
36.9 8.7 8.4 1.2
Ravindrarajah and Tuck (1994)[5]
36.4 5.6 5.4 0.8
33 11.8 11.6 1.6
34.2 10.9 11.9 1.5
36.4 10.2 1.53
29 15 2.04
21 14 2.16
34.4 8.4 2.1
22.2 15.2 2.23
Babu and Babu (2003) [7] 22.2 17.6 2.2
33.2 12.4 2.15
28.7 11.6 1.83
28.7 14.2 2.02
26.3 15.6 2.22
26 19.8 2.32
25 22.1 2.31
25 21.4 2.62
40 17.6 2.14
40 16.7 2.53
55 10.6 1.32
55 9.9 2.08
Chen and Liu (2004) [9]
25 25.7 2.4
25 25.9 2.73
40 20.1 2.22
40 20.8 2.68
55 11.3 1.73
55 10.9 2.31
Miled et al. (2004) [10] 8.1 5.633
66.5 1.1 0.64
58 2.3 2.1 0.89
49 3.83 4.33 1.15
Babu et al. (2005) [12]
38 6 8.45 1.4
28.5 7.8 9 2.04
16.3 12.5 16 2.34
14 2.4
16.5 3.8
Chen and Liu (2005) [13] 17 4.5
15.2 4.6
12.2 3.9

73
Chapter 3: Hardened concrete properties

Table 3.8: Mechanical properties of LWC containing EPS beads from experiments in literature (continued)

EPS ࢌᇱࢉ ࡱࢉ ࢌ࢘ ࢌࢉ࢚


References
Vol.% (MPa) (GPa) (MPa) (MPa)
16.8
16.3
15.5
Chen and Liu (2005) [13] 13.6
11.2
22.6
20.7
50 7 1.4
35 11 1.65
20 20 2.7
Babu et al. (2006) [14]
50 5 0.87
35 9 1.45
20 17 2.87
16.5 3.8
17 4.5
15.2 4.6
12.2 3.9
16.3 3.4
15.5 3.6
13.6 3.5
11.2 3.3
20.7 4
19.2 3.9
17 3.9
14.7 3.3
Chen and Liu (2007) [17]
24.3 4.2
25.4 4.3
20.4 4.4
18.8 4.2
20.3 4.8
18.6 5.2
15.5 4.9
13.4 4.2
20.3 3.9
17 4.3
16.2 4.4
13.8 4.2
15 42 4.5
35 21.5 3
55 8.5 1.55
15 44 4.9
35 23.5 3.3
55 9.5 1.8
Chen and Fang (2009) [22]
15 46 5
35 24.5 3.5
55 10 2
15 46.8 5.3
35 25 3.55
55 10.5 2.1
19.4 40 4.34
Chen and Fang (2011) [24] 19.4 48 4.89
19.4 31 3.67
74
Chapter 3: Hardened concrete properties

Table 3.8: Mechanical properties of LWC containing EPS beads from experiments in literature (continued)

EPS ࢌᇱࢉ ࡱࢉ ࢌ࢘ ࢌࢉ࢚


References
Vol.% (MPa) (GPa) (MPa) (MPa)
19.4 37 4.12
19.4 26 3.27
19.4 33 3.82
37.6 18 2.57
37.6 22 2.93
37.6 12 1.97
37.6 15 2.28
Chen and Fang (2011) [24] 37.6 10 1.75
37.6 13 2.08
55.8 6 1.25
55.8 8 1.51
55.8 4 0.96
55.8 5 1.11
55.8 3.5 0.88
55.8 4.5 1.04
10 26.58
15 28.86
22.5 31.14
30 22.78
10 24.30
15 25.82
Madandoust et al. (2011) [26]
22.5 18.99
30 20.51
10 22.03
15 17.47
22.5 18.23
30 19.75
15 33 3.30 3.94
25 16.7 2.26 2.71
40 9.8 2.39 2.82
55 5.4 2.52 3.12
15 35 1.57 2.35
25 24.4 2.04 2.94
Sadrmomtazi et al. (2012) [28]
40 10.2 1.26 1.71
55 2.9 1.30 1.82
15 22.4 0.96 1.35
25 10.6 1.00 0.94
40 7.1 0.70 0.82
55 3.5 0.74 0.76
15 20.77
20 17.79
25 11.22
25 17.86
Xu et al. (2012) [29] 15 7.85
20 15.68
20 13.52
25 18.56
15 12.98
13 25 17.84
Trussoni et al. (2013) [35]
12.4 23.25 15.33
5 1.76 0.23
10 2.07 0.17
15 2.76 0.19
20 2.37 0.15
Chen and Liu (2013) [37]
5 10.5 1.42
10 13 1.44
15 12.5 1.37
20 12 1.35

75
Chapter 3: Hardened concrete properties

Table 3.8: Mechanical properties of LWC containing EPS beads from experiments in literature (continued)

EPS ࢌᇱࢉ ࡱࢉ ࢌ࢘ ࢌࢉ࢚


References
Vol.% (MPa) (GPa) (MPa) (MPa)
60 11
100 5
60 8
Herki et al. (2013) [40]
100 3.7
60 5.5
100 3
15 kg/m3 33
25 16.7
40 9.8
15 27.8
Sadrmomtazi et al. (2013) [41] 25 24.4
40 10.2
15 22.4
25 10.6
40 6.7
10 18.12
10 18.87
10 19.63
20 16.23
Hidayat and Gunaedi (2014) [43] 20 16.99
20 17.55
30 14.91
30 15.66
30 16.04
0.49kg/m3 7.73 6.6 2.52
0.51 2.53 0.82
Kekanović et al. (2014) [44]
0.53 6.2 0.39
0.53 6.7
2.64kg/m3 46.3 9.11 6.85
5.76 23.6 6.87 5.37
8.75 20.3 5.79 4.24
10.9 12.2 3.72 2.4
Liu and Chen (2014) [45]
2.47 37.8 6.21 5.77
5.18 17.4 5.4 4.53
7.8 11.6 3.99 2.77
9.59 7.5 3.39 1.38
13.2kg/m3 5.5
Mousavi et al. (2014) [46]
7 6.3
55 16
56 14
Schackow et al. (2014) [47] 58 12
60 10
64 8
5 33.75
10 28.75
15 26.7 2.8
Tamut et al. (2014) [48]
20 22.9
25 20.4
30 17.1 2.4

76
Chapter 3: Hardened concrete properties

Table 3.8: Mechanical properties of LWC containing EPS beads from experiments in literature (continued)

EPS ࢌᇱࢉ ࡱࢉ ࢌ࢘ ࢌࢉ࢚


References
Vol.% (MPa) (GPa) (MPa) (MPa)
50 10.3
50 15.54
Nekooie et al. (2015) [49]
50 14.41
50 14.07
5 33 29.71 3.1
10 25.9 24.46 2.8
15 16.8 20.5 2.2
20 12.5 14.47 2.1
25 9.1 12.49 1.3
30 3.8 8.71 1.1
35 3.6 6.15 0.9
40 1.9 3.6 0.7
Xu et al. (2015) [50]
5 28.6 27.36 2.8
10 22.3 23.4 2.6
15 18 21.36 2.3
20 14.8 14.95 1.6
25 12.5 13.73 1.6
30 10.7 11.29 1.5
35 8.8 9.2 1.1
40 5.3 6.86 1
20 37 4.2
Li et al. (2015) [51] 40 22 3
60 11 1.87
5.32kg/m3 31 19.89
Pecce et al. (2015) [53]
5.09 20.5 18.55
10 23 2.2
15 20 2.02
22.5 16 1.81
30 15 1.77
Ranjbar and Mousavi (2015) [54]
10 31 2.33
15 25 2.23
22.5 22 2.14
30 19 1.95

3.4.1. Modulus of Elasticity

3.4.1.1. Experimental and analytical database for modulus of elasticity

Tables (3.4) to (3.8) present a comprehensive summary of the previously conducted


experiments in the literature. They show the components and proportions of the mixtures
involved in making LWC containing EPS beads. Different types of chemical and mineral
admixture are included in the mixtures. Also different methods have been applied to cure the
samples before testing. Various types of cementitious materials and natural and artificial
aggregates in the mixtures give different ranges of mechanical properties for hardened concrete.

The main properties of hardened concrete investigated in the present study include
compressive strength, modulus of elasticity, modulus of rupture, splitting tensile strength and

77
Chapter 3: Hardened concrete properties

stress-strain curve in the EPS-LWC mixtures. However, a wide range of existing models for
mechanical properties of conventional concrete and self-compacting concrete have been
collected from the existing literature. These data are compared and evaluate with the outcomes
of this study.

A general comparison between the mechanical properties of LWC made with different types
of lightweight aggregates and the conventional concrete is presented in Chapter Two. In this
chapter, mechanical properties of LWC containing EPS beads are investigated. New models are
developed for different properties of hardened concrete and evaluated by the existing models.

Considering the modulus of elasticity, Table (3.8) gives a considerable amount of


experimental data of EPS-LWC. In this section, the proposed model to predict the modulus of
elasticity (Ec) of EPS-LWC is based on the regression analysis of the experimental data. The
power type models in most of the existing codes and empirical equations are selected as the
basis of the newly developed models in this study.

3.4.1.2. Proposed model of modulus of elasticity in this study

An important part of the experimental datasets for EPS-LWC is classified as non-structural


LWC. Hence, the models are proposed for three groups of non-structural EPS-LWC with the
compressive strength not exceeding 17 MPa, structural LWC with compressive strength above
17 MPa and finally LWC containing all ranges of the compressive strength. According to the
collected experimental data, the upper limit of the compressive strength of the proposed models
is different from the various mechanical properties of LWC mixtures that are mentioned in the
proposed models.

Equation (3.1) is a general form of the proposed model to predict MoE from compressive
strength. The proposed model includes different types of natural and artificial aggregates, fillers
and chemical admixtures, different size and content of EPS beads along with the variable water
to binder ratios in the mixtures. Table (3.9) gives the specialized coefficients of the proposed
model for each group of the EPS-LWC. To evaluate the accuracy of proposed model,
correlation factor (R2) in the least square method of regression analysis is given in the table.

‫ܧ‬௖ ൌ ߙǤ ሺ݂௖ᇱ ሻఉ  Eq. (3. 1)

Where; Ec is modulus of elasticity (GPa), ݂௖ᇱ is compressive strength (MPa) and α, β are
coefficients.

78
Chapter 3: Hardened concrete properties

Table 3.9: Coefficients of equation (3.1) and the compressive strength limit

α β R2 ܎‫܋‬ᇱ limit (MPa)


Non-Structural EPS-LWC 0.9101 1.026 0.93 ≤ 17
Structural EPS-LWC 4.21 0.508 0.22 >17
EPS-LWC 1.4447 0.8511 0.89 3-35

Figure (3.1) shows the predicted versus experimental values of modulus of elasticity in two
ranges of compressive strength. As shown, the proposed equations accurately predict the
modulus of elasticity of EPS-LWC, especially for non-structural concrete.

Figure 3.1: Comparison of the experimental data with the predicted modulus of elasticity in EPS-LWC

3.4.1.3. Comparison of the analytical models of modulus of elasticity

In absence of accurate testing measurements, there are several analytical and numerical
models in the literature used to predict the modulus of elasticity of different types of concrete
mixtures. Tables (3.10) show some of these existing models to estimate the MoE of LWC.
These models vary in complexity and precision in the calculations.

Figure (3.2) shows the MoE versus compressive strength in LWC. Comparison of the
proposed models in Table (3.9) for EPS-LWC in non-structural data and all data ranges of
compressive strength are also presented.

This study utilizes EPS-LWC with a density of 2000 kg/m3 to manufacture the testing
specimens and one-way slabs for instantaneous and long-term deflection. Therefore, the
comparison is performed for that specified density in Figure (3.2).

As shown in Figure (3.2), the models in selected codes of practice give a comparable
prediction for MoE in LWC in all ranges of compressive strength. The empirical model
79
Chapter 3: Hardened concrete properties

proposed by Min and Gjorv (1991) significantly underestimates the MoE in LWC with the
density of 2000 kg/m3. The proposed models in this study underestimate MoE in the range of
compressive strength below 17 MPa (non-structural LWC) and over-estimate the MoE for
compressive strength above 17 MPa. The differences in overestimated MoE values by the
proposed models are considerable.

Since the density of LWC is a major parameter in the prediction of MoE, Figure (3.3)
compares predictions of the models in most frequently used codes of practice for the lower limit
of density (1280 kg/m3). Comparing Figures (3.2) and (3.3) show the variation of MoE values
by changing the density in LWC. Even, when considering similar prediction of the selected
codes for LWC with density of 2000 kg/m3, the same models estimate different MoE values for
LWC with density of 1280 kg/m3.

Table 3.10: Existing models to predict the modulus of elasticity of lightweight concrete

Reference ࡱࢉ ൌ ࡭࢝࡮ Ǥ ሺࢌᇱࢉ ሻ࡯ ൅ ࡰ : Ec(GPa), ࢌᇱࢉ (Mpa) Notes


‫ܧ‬௖ ൌ ͲǤͲͶ͵݇ଵ ߛ ଵǤହ ሺ݂௖ᇱ ሻ଴Ǥହ 3
kg/m and MPa 1280< γ < 2400 kg/m3
AASHTO LRFD (2007) [57]
K1 = 1.0 unless determined by physical test 14≤ ݂௖ᇱ ≤ 48 MPa
1440<γ< 2800 kg/m3
Rizkalla (NCHRP)(2007) [58] ‫ܧ‬௖ ൌ ͵ͳͲͲͲͲ݇ͳߛ ଶǤହ Ǥ ሺ݂௖ᇱ ሻ଴Ǥଷଷ ݂௖ᇱ < 165 MPa
Ec < 74 GPa
ACI 318- 08 [59] ‫ܧ‬௖ ൌ ͲǤͲͶ͵‫ݓ‬ଵǤହ ሺ݂௖ᇱ ሻ଴Ǥହ MPa, GPa, kg/m 3
for 1440< γ <2500 kg/m3
Modified ACI 318-95 [60] ‫ܧ‬௖ ൌ ͵ͲǤͳ͸ߛଵǤହ ሺ݂௖ᇱ ሻ଴Ǥହ ൅ ͶͺͶʹͲͲ Ec(psi), ݂௖ᇱ (psi), w(pcf)
‫ܧ‬௖ ൌ ͲǤͲͶ͵ߛଵǤହ ሺ݂௖ᇱ ሻ଴Ǥହ MPa, kg/m3
ACI-209-2R-08 [61] 1440< γ <2500 kg/m3
‫ܧ‬௖ ൌ Ͷ͹͵Ͳሺ݂௖ᇱ ሻ଴Ǥହ MPa
‫ܧ‬௖ ൌ ͳͶͲͲͲ ൅ ͵ʹͷͲሺ݂௖ᇱ ሻ଴Ǥହ
ACI 312-92 [62] ‫ܧ‬௖ ൌ ͻͷͲͲሺ݂௖ᇱ ൅ ͺሻଵȀଷ GPa, MPa
‫ܧ‬௖ ൌ ͲǤͲͶ͵ߛଵǤହ ሺ݂௖ᇱ ሻ଴Ǥହ
ߛ
‫ܧ‬௖ ൌ ሺ͵͵ͲͲሺ݂௖ᇱ ሻ଴Ǥହ ൅ ͸ͻͲͲሻሺ ൗʹ͵ͲͲሻଵǤହ 20 ≤ ݂௖ᇱ ≤ 40 MPa
CSA A23.3-04 [63]
‫ܧ‬௖ ൌ ͶͷͲͲሺ݂௖ᇱ ሻ଴Ǥହ 1500< γ < 2500 kg/m3

NZS-3101-95 [64] ‫ܧ‬௖ ൌ ሺ͵͵ʹͲሺ݂௖ᇱ ሻ଴Ǥହ ൅ ͸ͻͲͲሻሺ‫ݓ‬ൗʹ͵ͲͲሻଵǤହ Ec in MPa


‫ܧ‬௖ ൌ ͲǤͲͶ͵ߛଵǤହ ሺ݂௖ᇱ ሻ଴Ǥହ ݂௖ᇱ <40 MPa
AS-3600-2900 [65]
‫ܧ‬௖ ൌ ͷͲͷͲሺ݂௖ᇱ ሻ଴Ǥହ normal weight concrete 0< Secant < 0.4 fcm

‫ܧ‬௖ ൌ ͻǤͷሺ݂௖ᇱ ሻ଴Ǥଷ ሺ ሻଵǤହ GPa, kg/m3
ଶସ଴଴
NS-3473 (1992) [66]
‫ܧ‬௖ ൌ ͻͷͲͲሺ݂௖ᇱ ሻ଴Ǥଷ

‫ܧ‬௖ ൌ ͻǤͷሺ݂௖ᇱ ሻ଴Ǥଷ ሺ ሻଵǤହ GPa, kg/m3
ଶସ଴଴
NS-3473 (1992) [66]
‫ܧ‬௖ ൌ ͻͷͲͲሺ݂௖ᇱ ሻ଴Ǥଷ

80
Chapter 3: Hardened concrete properties

Table 3.10: Existing models to predict the modulus of elasticity of lightweight concrete (continued)

Reference ࡱࢉ ൌ ࡭࢝࡮ Ǥ ሺࢌᇱࢉ ሻ࡯ ൅ ࡰ : Ec(GPa), ࢌᇱࢉ (Mpa) Notes


Architecture institute of Japan (AIJ) ߛ ݂ᇱ
‫ܧ‬௖ ൌ ʹǤͳ ൈ ͳͲହ ሺ ൗʹǤ͵ሻଵǤହ ሺ ௖ ൗʹͲͲሻ଴Ǥହ Ec, ݂௖ᇱ : kgf/cm2 , γ : t/m3
[67]
Yusuf, I T; Jimoh, Y A (2013) [68] ‫ܧ‬௖ ൌ ͲǤͲͲͳሺ݂௖ᇱ ሻଵǤହସଽ
‫ܧ‬௖ ൌ Ͷ͹ͲͲሺ݂௖ᇱ ሻ଴Ǥହ  (ACI-318-08)
JSCE (2007) [69] 18 ≤ ݂௖ᇱ ≤80 MPa
‫ܧ‬௖ ൌ ͳͲǤ͹ͻʹ Žሺ݂௖ᇱ ሻ െ ͻǤͲ͸͹ͷ best fit
ߛ
RakMK-D3-2012 [70] ‫ܧ‬௖ ൌ ͷͲͲͲ൫ ൗʹͶͲͲ൯ሺ݂௖ᇱ ሻ଴Ǥହ

‫ܧ‬௖ ൌ ͲǤͲͶ͵ߛଵǤହ ሺ݂௖ᇱ ሻ଴Ǥହ MPa, kg/m3


NTE E.060(2009) [71]
‫ܧ‬௖ ൌ Ͷ͹͵Ͳሺ݂௖ᇱ ሻ଴Ǥହ MPa
BS 8110 (1997) [72] ‫ܧ‬௖ ൌ ሺʹͲ ൅ ͲǤʹ݂௖ᇱ ሻሺߛȀʹͶͲͲሻଶ GPa 20 ≤ ݂௖ᇱ ≤60 MPa
BS 5400-4(1990)[73] ‫ܧ‬௖ ൌ ͺǤ͸Ͷ͹ͷሺ݂௖ᇱ ሻ଴Ǥଷସ଼  20 ≤ ݂௖ᇱ ≤60 MPa
SP 52-101-2003 [74] ‫ܧ‬஼ ൌ ͳͳǤ͸ʹͷ Žሺ݂௖ᇱ ሻ െ ͹ǤͶ͹ͳ͵ 10 ≤ ݂௖ᇱ ≤ 60 MPa
‫ܧ‬௖ ൌ ͵Ǥ͵ʹͲሺ݂௖ᇱ ሻ଴Ǥହ ൅ ͸ǤͻͲͲሻሺ‫ݓ‬ൗʹ͵Ͷ͸ሻ
Carrasquillo, et al.(1981) [75] GPa , kg/m3
‫ܧ‬௖ ൌ ͵Ǥ͵ʹͲሺ݂௖ᇱ ሻ଴Ǥହ ൅ ͸ǤͻͲͲ
Best fit with:
ߛ ଵǤହ
‫ܧ‬௖ ൌ ൫ ൗʹ͵ͲͲ൯ ሺ͵ͲͲͲሺ݂௖ᇱ ሻ଴Ǥହ ൅ ͸ͻͲͲሻሻ - Eq. 8-6 in CSA-
Soleymani (2006) [76]
ߛ ଵǤହ A23.3(1994)
‫ܧ‬௖ ൌ ൫ ൗʹ͵ͲͲ൯ ሺ͵͵ͲͲሺ݂௖ᇱ ሻ଴Ǥହ ൅ ͸ͻͲͲሻ
- Eq. 8-6 in CSA-23.3(1994)
Haranki (2009) [77] ‫ܧ‬௖ ൌ ͵ͳǤͻʹ ൈ ߛ ଵǤହ
ሺ݂௖ᇱ ሻ଴Ǥହ ൅ ͵Ͷͷǡ͵ʹͺ psi , lb/ft3
Ahmad and Shah (1985) [78] ‫ܧ‬௖ ൌ ͵Ǥ͵ͺ ൈ ͳͲିହ ൈ ߛ ଶǤହ ሺඥ݂௖ᇱ ሻ଴Ǥ଺ହ

Jobse and Mustafa (1984) [79] ‫ܧ‬௖ ൌ ͲǤͳͲ͵ߛ ଵǤହ ሺඥ݂௖ᇱ ሻ଴Ǥହ
Cook (1989) [80] ‫ܧ‬௖ ൌ ͵Ǥʹʹ ൈ ͳͲିହ ൈ ߛ ଶǤହ ሺ݂௖ᇱ ሻ଴Ǥଷଵହ kg/m3 , MPa
Min and Gjorv (1991) [81] ‫ܧ‬௖ ൌ ͳǤͳͻሺ݂௖ᇱ ሻଶȀଷ GPa , MPa For light-weigth concrete
1280< γ <2400 kg/m3
Jensen (1943) [82] ‫ܧ‬௖ ൌ ͸ ൈ ͳͲ଺ Ȁሺͳ ൅ ʹͲͲͲȀ݂௖ᇱ ሻ psi
14≤ ݂௖ᇱ ≤48 MPa
Pauw (1960) [83] ‫ܧ‬௖ ൌ ͳ͵Ǥͺʹߛ ଵǤ଻ଽ ሺ݂௖ᇱ ሻ଴Ǥସସ
ߛ ଵǤହ
Slate et al. (1986) [84] ‫ܧ‬௖ ൌ ሺ͵͵ʹͲሺ݂௖ᇱ ሻ଴Ǥହ ൅ ͸ͺͻͷሻ൫ ൗʹ͵ʹͲ൯

Ravindrarajah and Tuck (1994) [5] ‫ܧ‬௖ ൌ ͳǤͳͶ͸ሺ݂௖ᇱ ሻ଴Ǥହ


ଵǤହ
ߛ
Khan (1990) ‫ܧ‬௖ ൌ ሺ͵Ǥ͵ʹሺ݂௖ᇱ ሻ଴Ǥହ ൅ ͸Ǥͻሻቀ ൗʹ͵ʹͷቁ MPa
ଵǤହ
ߛ
Morales (1982) [85] ‫ܧ‬௖ ൌ ሺͶͲͲͲͲሺ݂௖ᇱ ሻ଴Ǥହ ൅ ͳͲͲͲͲͲͲሻቀ ൗͳͶͷቁ psi
଴Ǥଽ
Meyer (2002) [86] ‫ܧ‬௖ ൌ ሺ͵͵ͲͲͲሺ݂௖ᇱ ሻ଴Ǥହ ൅ ͶͲͲͲͲͲͲሻ൫‫ݓ‬ൗʹͶʹ൯ psi
ߛ ଶ
Alengaram et al. (2011) [87] ‫ܧ‬௖ ൌ ͷሺ݂௖ᇱ ሻଵȀଷ ൫ ൗʹͶͲͲ൯

Ahmmad et al. (2014) [88] ‫ܧ‬௖ ൌ ͲǤͲͲʹ͸ሺ݂௖ᇱ ሻଶǤହହସ଺ GPa


Dinakar (2012) [89] ‫ܧ‬௖ ൌ ͶǤʹሺ݂௖ᇱ ሻ଴Ǥହ GPa
଴Ǥହ ଵǤହ
Tassw and Lubell (2012) [90] ‫ܧ‬௖ ൌ ͲǤͲ͵ͳ݂௖ᇱ ߛ

The presented study investigates the mechanical properties of lightweight concrete. Also, the
short-term and long-term behaviour of lightweight concrete slabs are evaluated

81
Chapter 3: Hardened concrete properties

comprehensively. To better analysis of the mechanical properties, it is tried to collect all the
published models in the articles and the technical reports in literature.

In total, 37 models to estimate the modulus of elasticity from compressive strength in


lightweight concrete are collected. The oldest model is developed in 1943 for the wide range of
density and compressive strength of lightweight concrete. Considering the lower density of
lightweight concrete in comparison with conventional concrete, the density limit is almost
applied in all the models. Also, limitations on the compressive strength are applied in most of
the models. Some models are also applicable for conventional concrete by changing the density
related coefficients.

Figure 3.2: Comparison of the existing MoE models with the proposed models (Eq.3.1) (density 2000 kg/m3)

82
Chapter 3: Hardened concrete properties

Figure 3.3: Comparison of the existing MoE models with the proposed models (Eq. 3.1) (density 1280 kg/m3)

3.4.2. Modulus of Rupture in Lightweight Concrete

3.4.2.1. Experimental and analytical database in EPS-LWC

Considering the modulus of elasticity, Table (3.8) presents a large number of experimental
data relating to EPS-LWC. In this section, the proposed model to predict the Modulus of
Rupture (MoR) of EPS-LWC is based on regression analysis of the experimental data. The
power type models in most of the existing codes of practice and empirical equations are selected
as the basis of the newly developed models in this study.

3.4.2.2. Proposed analytical model in this study

From the experimental results, some equations have been developed to predict the modulus of
rupture from compressive strength. According to Chapter 2, the compressive strength above 17
MPa is considered as a minimum strength requirement of structural LWC. Hence, the developed
equations are presented for classified ranges and the whole range of compressive strength
respectively.

Equation (3.2) is a general form of the proposed model of predicting MOR from compressive
strength. The proposed models include different types of natural and artificial aggregates, fillers
and chemical admixtures, different size and content of EPS beads along with the variable ratios
of water to binder in the mixtures. Table (3.11) gives the specialized coefficients of the
proposed model for each group of the EPS-LWC. To evaluate the proposed models, correlation
factor (R2) in the least square method of regression analysis is given in the table as well.
83
Chapter 3: Hardened concrete properties

݂௥ ൌ ߮Ǥ ሺ݂௖ᇱ ሻ௽ Eq. (3. 2)

Where; fr is modulus of rupture (MPa), ݂௖ᇱ is compressive strength (MPa) and φ, ߎ are coefficients.

Table 3.11: Coefficients of Equation (3.2) and the compressive strength limit

φ ߎ R2 Range of ࢌᇱࢉ (MPa)


Non-Structural EPS-LWC 0.524 0.7398 0.91 ≤ 17
Structural EPS-LWC 1.266 0.4062 0.52 >17
EPS-LWC 0.5769 0.6756 0.83 3-40

Figure (3.4) shows the predicted versus experimental values of MoR of EPS-LWC in two
ranges of compressive strength. As shown, the developed equations accurately predict MoR of
EPS-LWC, especially for non-structural concrete. However, more experimental data is required
to develop accurate models for structural parts.

Figure 3.4: Comparison of the modulus of rupture in experimental data and prediction of Eq. (3.2)

3.4.3. Comparison of the Analytical Models of Modulus of Rupture

Modulus of rupture is predicted from compressive strength in various types of empirical


equations and the models in codes of practice for different types of concrete. Table (3.13) gives
some existing models in literature to estimate the MOR of LWC in the absence of accurate
testing methods. The calculations of these models vary in complexity and precision.

Figure (3.3) shows the MoR versus compressive strength of LWC and compares the proposed
models in Table (3.12) for EPS-LWC in non-structural data range and all data ranges of the
compressive strength.

84
Chapter 3: Hardened concrete properties

Table 3.12: Existing models to predict the modulus of rupture of lightweight concrete

Reference ࢌ࢘ ൌ ࡭ሺࢌᇱࢉ ሻ࡮ ൅ ࡯ Notes


Slate et al. (1986) [84] ݂௥ ൌ ͲǤ͵ͷሺ݂௖ᇱ ሻ଴Ǥହ ʹͲ ൑ ݂௖ᇱ ൑ ͸Ͳ MPa
Yusuf, I T; Jimoh, Y A (2013) [68] ݂௥ ൌ ͲǤ͵ͻͻሺ݂௖ᇱ ሻ଴Ǥହଽଵ
Cousins et al. (2013)[91] ݂௥ ൌ ݇ሺ݂௖ᇱ ሻ଴Ǥ଺ସ  ͲǤʹ͸ ൑ ݇ ൑ ͲǤ͵͸
Alengaram et al. (2013)[92] ݂௥ ൌ ͲǤ͵ሺ݂௖ᇱ ሻଶȀଷ
଴Ǥହ
ACI 213R-03 [93] ݂௥ ൌ ͲǤͶͷ݂௖ᇱ
଴Ǥହ
Tassew and Lubell (2012) [90] ݂௥ ൌ ͲǤ͵͵݂௖ᇱ
଴Ǥହ
ACI 318-95 [60] ݂௥ ൌ ͲǤͶ͸݂௖ᇱ
଴Ǥହ
FHWA (2013) [94] ݂௥ ൌ ͲǤͶ͸݂௖ᇱ
଴Ǥହ
݂௥ ൌ ͲǤͶ͸ͷ݂௖ᇱ All LWC
ACI-318-08 [59]
଴Ǥହ Sand LWC
݂௥ ൌ ͲǤͷʹ͹݂௖ᇱ
଴Ǥହ
AASHTO (2012) [95] ݂௥ ൌ ͲǤͳ͹݂௖ᇱ
଴Ǥହ
Bogas (2014) [96] ݂௥ ൌ ͲǤ͸ͻ݂௖ᇱ

The total number of 11 models since 1986 in the codes of practice and the published research
articles and technical reports are collected. The models are used to estimate the modulus of
rupture from compressive strength in lightweight concrete. Majority of the models apply a
power type relationship to estimate the modulus of rupture from compressive strength.
However, effect of the concrete density is considered in some models. Also, limitation on the
compressive strength is applied in some models.

Figure 3.5: Comparison of the existing MoR models with the proposed models (density 1280 kg/m3)

85
Chapter 3: Hardened concrete properties

From Figure (3.5), it is evident that the proposed model for non-structural EPS-LWC and
EPS-LWC in all ranges of the compressive strength, predict higher values of MoR than other
models presented in Table (3.9). Among the existing models, AASHTO (2013) and Slate (1986)
models give the lowest predictions, while the models presented in ACI-213-03, FHWA (2013),
Cousins et al. (2013) and Alengaram et al. (2013) predict close values of MoR for structural and non-
structural LWC.

3.4.4. Splitting Tensile Strength in Lightweight Concrete

3.4.4.1. Experimental and analytical database in EPS-LWC

Considering the modulus of elasticity, Table (3.5) gives a considerable amount of


experimental data in regard to EPS-LWC. In this section, the proposed model of predicting the
Splitting Tensile Strength (STS) of EPS-LWC is based on the regression analysis of the
experimental data. The power type relationships in most of the existing codes and empirical
equations are selected as the basis of developing new models.

3.4.4.2. Proposed analytical model in this study

From the experimental results some equations are developed to predict STS of LWC from
compressive strength. The developed equations are presented for classified ranges of
compressive strength.

Equation (3.3) is a general form of the proposed model to predict STS from compressive
strength. The proposed models include different types of natural and artificial aggregates, fillers
and chemical admixtures, different size and content of EPS beads, along with the variable ratios
of water to binder in the mixtures. Table (3.13) gives the specialized coefficients of the
proposed model for each group of the EPS-LWC. To evaluate the proposed models, correlation
factor (R2) in the least square method is given in the table also.

݂௖௧ ൌ ߮Ǥ ሺ݂௖ᇱ ሻఠ  Eq. (3. 3)

Where; fct is splitting tensile strength (MPa), ݂௖ᇱ is compressive strength (MPa) and φ, ω are
coefficients.

Table 3.13: Coefficients of equation (3.3) and the compressive strength limit

φ ω R2 ܎‫܋‬ᇱ limit (MPa)


Non-Structural EPS-LWC 0.3705 0.6644 0.98 ≤ 17
Structural EPS-LWC 0.1862 0.8655 0.94 >17
EPS-LWC 0.2818 0.7378 0.811 1-50

86
Chapter 3: Hardened concrete properties

Figure (3.6) shows the predicted via experimental values of splitting tensile strength into two
distinct ranges of compressive strength. As shown, the developed equations accurately predict
the splitting tensile strength of EPS-LWC, especially for non-structural concrete.

(a) ࢌᇱࢉ ൏ ૚ૠࡹࡼࢇ

(b) ࢌᇱࢉ ൒ ૚ૠࡹࡼࢇ


Figure 3.6: Comparison of the experimental and predicted splitting tensile strength in EPS-LWC

3.4.4.3. Comparison of the STS analytical models

Splitting tensile strength for different types of concrete are predicted based on the
compressive strength in various types of empirical equations and the models in some codes of
practice. Table (3.10) gives some existing models in literature to estimate the STS of LWC in
the absence of accurate testing methods. These models vary in complexity and precision in the
calculations.
87
Chapter 3: Hardened concrete properties

Figure (3.7) shows the STS versus compressive strength of LWC, and compares the proposed
models in Table (3.14) in non-structural EPS-LWC and all data ranges of compressive strength
in EPS-LWC.

Table 3.14: Existing models to predict the splitting tensile strength of lightweight concrete

Reference ࢌࢉ࢚ ൌ ࡭ሺࢌᇱࢉ ሻ࡮ ൅ ࡯ Notes


Shafigh et al. (2013) [97] ݂௖௧ ൌ ͲǤʹ͵ሺ݂௖ᇱ ሻ଴Ǥ଺ସ
Cousins et al. (2013)[91] ݂௖௧ ൌ ͲǤʹͳሺ݂௖ᇱ ሻ଴Ǥହ
Dinakar (2013) [89] ݂௖௧ ൌ ͲǤʹ͸ͺሺ݂௖ᇱ ሻ଴Ǥ଺଻ହ
Slate et al. (1986) [84] ݂௖௧ ൌ ͲǤͶͳሺ݂௖ᇱ ሻ଴Ǥହ ʹͳ ൑ ݂௖ᇱ ൑ ͸ʹ‫ܽ݌ܯ‬
ͲǤ͸ߛ
EC-2 (2002) [98] ݂௖௧ ൌ ͲǤ͵͵ሺ݂௖ᇱ ሻ଴Ǥ଺଻ ሺͲǤͶ ൅ ሻ ݂௖ᇱ ൒ ͷͲ‫ܽܲܯ‬
ʹʹͲͲ
Wegian (2012) [99] ݂௖௧ ൌ ͲǤͲͺ͹݂௖ᇱ െ ͲǤͺ͹
Bogas (2014) [96] ݂௖௧ ൌ ͲǤͶ͹ሺ݂௖ᇱ ሻ଴Ǥହ
Yew et al. (2014) [100] ݂௖௧ ൌ ͲǤͶ͹ሺ݂௖ᇱ ሻ଴Ǥହ
݂௖ᇱ ൌ ͳͷǤ͵͸݂௖௧ െ ͸Ǥͻ͵
Kou et al. (2009) [101] ʹͲ ൑ ݂௖ᇱ ൑ Ͷͳ
݂௖௧ ൌ ͲǤʹͶሺ݂௖ᇱ ሻ଴Ǥ଺ଽ

To estimate the splitting tensile strength from compressive strength, totally 9 models in the
design codes and empirical equations since 1986 are presented. The power type relationship is
used in majority of the models. Also, the compressive strength is limited in some models.
Density is the only other parameter included in the models.

Figure (3.7) shows the splitting tensile strength versus compressive strength of EPS-LWC. It
compares the proposed models of EPS-LWC with existing models presented in Table (3.14).

According to Figure (3.7), it is evident that the proposed model for the structural and non-
structural EPS-LWC, predicts higher values of STS than other models presented in Table (3.10).
The models presented by Cousins et al. (2013) and Wegian (2012) give the lowest estimation of
STS for LWC in whole ranges of compressive strength. All other models in Figure (3.7) have
comparable prediction for STS for structural and non-structural LWC.

88
Chapter 3: Hardened concrete properties

Figure 3.7: Comparison of the proposed and existing models of splitting tensile strength (density 2000 kg/m3)

3.4.5. Relationship between Modulus of Rupture and Modulus of Elasticity

The relationship between MoE and MoR in EPS-LWC is illustrated in Eq. (3.4). The related
coefficients and the ranges of applied compressive strength are presented in Table (3.15).

݂௥ ൌ ߩǤ ሺ‫ܧ‬௖ ሻఙ  Eq. (3. 4)

Where; fr is modulus of rupture (MPa), Ec is modulus of elasticity (GPa) and ρ, σ are


coefficients.

Table 3.15: Coefficients of Equation (3.4) and compressive strength limit

ρ σ R2 ܎‫܋‬ᇱ limit (MPa)


EPS-LWC 0.7027 0.5591 0.78 ≤ 50 MPa

3.4.6. Relationship between splitting tensile strength and modulus of elasticity

The relationship between MoE and STS in EPS-LWC is illustrated by Equation (3.5). The
relevant coefficients and the ranges of applied compressive strength are also presented in Table
(3.16).
݂௖௧ ൌ ߯Ǥ ሺ‫ܧ‬௖ ሻట Eq. (3.5)

Where; fct is splitting tensile strength (MPa), Ec is modulus of elasticity (GPa) and φ, ω are
coefficients.

89
Chapter 3: Hardened concrete properties

Table 3.16: Coefficients of Equation (3.5)

χ ψ R2 Range of ܎‫܋‬ᇱ
EPS-LWC 0.3867 0.5746 0.95 ≤ 50 MPa

3.5. MECHANICAL PROPERTIES OF CONVENTIONAL AND SELF-


COMPACTING CONCRETE

In the absence of accurate testing systems, the models proposed in international codes of
practice and the empirical models developed by some experimental investigations are useful
tools in estimating the mechanical properties of concrete from compressive strength.

Along with the models presented for LWC in the previous parts of this chapter, the models
for different properties of conventional concrete and self-compacting concrete are collected and
presented also.

Modulus of elasticity, modulus of rupture and splitting tensile strength are the basic
properties of concrete used to investigate structural behaviour. In addition, the compressive
stress-strain curve and creep and shrinkage are other properties of concrete that are studied in
the next sections.

3.5.1. Conventional Concrete

3.5.1.1. Modulus of elasticity

Table (3.17) shows the existing models used in some codes of practice to predict the modulus
of elasticity in conventional concrete.

Table (3.18) presents some empirically developed models to predict the modulus of elasticity
of conventional concrete. The experimental models are modified form of the models in the most
frequently used codes of practice worldwide.

90
Chapter 3: Hardened concrete properties

Table 3.17: Existing models of the MoE of conventional concrete in codes of practice

Design code Model Limits and coefficients


‫ܧ‬௖ ൌ ͲǤͲͶ͵ߛଵǤହ ሺ݂௖ᇱ ሻ଴Ǥହ MPa, kg/m 3
ACI 318- 08 [59] 1440< γ<2500 kg/m3
‫ܧ‬௖ ൌ Ͷ͹͵Ͳሺ݂௖ᇱ ሻ଴Ǥହ MPa
OHBDC(1983) [102] ‫ܧ‬௖ ൌ ͷͲͲͲሺ݂௖ᇱ ሻ଴Ǥହ psi
ߛ ݂ᇱ ‫ܧ‬௖ , ݂௖ᇱ : kgf/cm2 , γ = t/m3
AIJ (1985) [67] ‫ܧ‬௖ ൌ ʹǤͳ ൈ ͳͲହ ሺ ൗʹǤ͵ሻଵǤହ ሺ ௖ ൗʹͲͲሻ଴Ǥହ

α=1.2 basalt, dense limestone,


CEB-FIP (1990) [103] ‫ܧ‬௖ ൌ ʹͳͷͲͲߙሺ݂௖ᇱ ȀͳͲሻଵȀଷ 1= quartzite , 0.9 limestone,
0.7 sandstone aggregate

‫ܧ‬௖ ൌ ͻǤͷሺ݂௖ᇱ ሻ଴Ǥଷ ሺ ሻଵǤହ GPa, kg/m 3
ଶସ଴଴
NS-3473(1992) [66]
‫ܧ‬௖ ൌ ͻͷͲͲሺ݂௖ᇱ ሻ଴Ǥଷ
‫ܧ‬௖ ൌ ͳͶͲͲͲ ൅ ͵ʹͷͲሺ݂௖ᇱ ሻ଴Ǥହ

ACI 312-92 [62] ‫ܧ‬௖ ൌ ͻͷͲͲሺ݂௖ᇱ ൅ ͺሻଵȀଷ


‫ܧ‬௖ ൌ ͲǤͲͶ͵ߛଵǤହ ሺ݂௖ᇱ ሻ଴Ǥହ
CEB (1993) [104] ‫ܧ‬௖ ൌ ͳͲͲͲͲሺ݂௖ᇱ ൅ ͺሻଵȀଷ

GBJ 11-89 (1994) [105] ‫ܧ‬௖ ൌ ͳͲଶ Ȁ ቂʹǤʹ ൅ ቀ͵ͶǤ͹ൗ݂ ᇱ ቁቃ


modified ACI 318-95 [60] ‫ܧ‬௖ ൌ ͵ͲǤͳ͸ߛଵǤହ ሺ݂௖ᇱ ሻ଴Ǥହ ൅ ͶͺͶʹͲͲ Ec in psi; ݂௖ᇱ in psi; γ in pcf
ߛ
NZS-3101-95 [64] ‫ܧ‬௖ ൌ ሺ͵͵ʹͲሺ݂௖ᇱ ሻ଴Ǥହ ൅ ͸ͻͲͲሻሺ ൗʹ͵ͲͲሻଵǤହ ‫ܧ‬௖ ǡ ݂௖ᇱ : MPa
ACI 363R-92, (1997)
‫ܧ‬௖ ൌ ͵͵ʹͲሺ݂௖ᇱ ሻ଴Ǥହ ൅ ͸ͺͻͲ 21 < ݂௖ᇱ < 83 MPa
[106]
BS 8110 (1997) [72] ‫ܧ‬௖ ൌ ʹͲ ൅ ͲǤʹ݂௖ᇱ 20 ≤ ݂௖ᇱ ≤60 MPa

EHE (1998) [107] ‫ܧ‬௖ ൌ ͳͲǡͲͲͲඥ݂௖ᇱ
BS 5400-4(1999) [73] ‫ܧ‬௖ ൌ ͺǤ͸Ͷ͹ͷሺ݂௖ᇱ ሻ଴Ǥଷସ଼  20 ≤ ݂௖ᇱ ≤60 MPa
FHWA (2000) [108] ‫ܧ‬௖ ൌ ͵ͺ͵͹ሺ݂௖ᇱ ሻ଴Ǥହ 28 ≤ ݂௖ᇱ ≤193 MPa
TS-500 (2000) [109] ‫ܧ‬௖ ൌ ͵Ǥʹͷሺ݂௖ᇱ ሻ଴Ǥହ ൅ ͳͶ
IS 456 (BIS, 2000) [110] ‫ܧ‬௖ ൌ ͷͲͲͲሺ݂௖ᇱ ሻ଴Ǥହ
GDC (2000) [111] ‫ܧ‬௖ ൌ ͶǤ͹͸ሺ݂௖ᇱ ሻ଴Ǥହ
EC2-02 (2002) [98] ‫ܧ‬௖ ൌ ʹʹͲͲͲሺ݂௖ᇱ ȀͳͲሻ଴Ǥଷ MPa Secant between 0 and 0.4fcm
K1 =1.0, K2 =90th percentile upper bound and
‫ܧ‬௖ ൌ ͵͵ͲͲͲ݇ଵ ݇ଶ ሺͲǤͳͶ ൅
NCHRP-2003 [112] the 10th percentile lower bound
݂௖ᇱ ȀͳͲͲͲሻଵǤହ Ǥ ሺ݂௖ᇱ ሻ଴Ǥହ ksi
2320< γ <2480 kg/m3
NBR-6118 (2003) [113] ‫ܧ‬௖ ൌ ͷ͸ͲͲሺ݂௖ᇱ ሻ଴Ǥହ
IDC 3274 (2003) [114] ‫ܧ‬௖ ൌ ͷǤ͹ሺ݂௖ᇱ ሻ଴Ǥହ

91
Chapter 3: Hardened concrete properties

Table 3.17: Existing models of the MoE of conventional concrete in codes of practice (continued)

Design code Model Limits and coefficients


SP 52-101(2003) [74] ‫ܧ‬௖ = 11.652ln(݂௖ᇱ ) - 7.4713 10 ≤ ݂௖ᇱ ≤ 60 MPa
‫ܧ‬௖ ൌ Ec in MPa
ሺ͵͵ͲͲሺ݂௖ᇱ ሻ଴Ǥହ ߛ
CSA A23.3-04 [63] ൅ ͸ͻͲͲሻሺ ൗʹ͵ͲͲሻଵǤହ 20 ≤ ݂௖ᇱ ≤40 MPa
‫ܧ‬௖ ൌ ͶͷͲͲሺ݂௖ᇱ ሻ଴Ǥହ 1500< γ <2500 kg/m3

K0 (GPa) = 17 (ferro quartzite);


20 ( Jukskei granite);
SABS-0100 (1992)
‫ܧ‬௖ ሺ‫ܽܲܩ‬ሻ ൌ ‫ܭ‬଴ ൅ ߙ݂௖௨ 29 (Eikenhof andesit
Modified (2006) [115]
α (GPa/MPa) = 0.4 (ferro quartzite);
0.2 ( Jukskei granite and Eikenhof andesit
‫ܧ‬௖ ൌ Ͷ͹ͲͲሺ݂௖ᇱ ሻ଴Ǥହ  (ACI-318-08)
JSCE (2007) [69] 18 ≤ ݂௖ᇱ ≤80 MPa
‫ܧ‬௖ = 10.792ln(f’c) - 9.0675 best fit
‫ܧ‬௖ ൌ ͲǤͲͶ͵݇ଵ ߛ ଵǤହ ሺ݂௖ᇱ ሻ଴Ǥହ kg/m3 and MPa
AASHTO-LRFD (2007)
K1 = 1.0 unless determined by physical 1280< γ <2400 kg/m3
[57]
test 14≤ ݂௖ᇱ ≤48 MPa
‫ܧ‬௖ ൌ ͲǤͲͶ͵ߛଵǤହ ሺ݂௖ᇱ ሻ଴Ǥହ MPa, kg/m3
ACI-209-2R-08 [61] for 1440< γ <2500 kg/m3
‫ܧ‬௖ ൌ Ͷ͹͵Ͳሺ݂௖ᇱ ሻ଴Ǥହ MPa
‫ܧ‬௖ ൌ ͲǤͲͶ͵ߛଵǤହ ሺ݂௖ᇱ ሻ଴Ǥହ MPa, kg/m3
NTE E.060(2009) [71]
‫ܧ‬௖ ൌ Ͷ͹͵Ͳሺ݂௖ᇱ ሻ଴Ǥହ MPa
‫ܧ‬௖ ൌ ͲǤͲͶ͵ߛଵǤହ ሺ݂௖ᇱ ሻ଴Ǥହ ݂௖ᇱ < 40 MPa,
AS-3600(2009) [65]
‫ܧ‬௖ ൌ ͷͲͷͲሺ݂௖ᇱ ሻ଴Ǥହ Secant between 0 and 0.4fcm
AS-3600(2009) [65] ‫ܧ‬௖ ൌ ߛଵǤହ ሺͲǤͲʹͶሺ݂௖ᇱ ሻ଴Ǥହ ൅ ͲǤͳʹሻ ݂௖ᇱ > 40 MPa
ߛ
RakMK-D3(2012) [70] ‫ܧ‬௖ ൌ ͷͲͲͲ൫ ൗʹͶͲͲ൯ሺ݂௖ᇱ ሻ଴Ǥହ

The total number of 53 models since 1943 in the codes of practice and the published research
articles and technical reports are collected to compare with the models for modulus of elasticity
in the lightweight concrete and self-compacting concrete. Majority of the models are the power
type relationship between the compressive strength and modulus of elasticity. Density of
concert, type of aggregates and the limitation on the applied compressive strength in the models
are the other effective parameters.

92
Chapter 3: Hardened concrete properties

Table 3.18: Empirical models to predict the modulus of elasticity of conventional concrete

Researchers Model Limits and coefficients


1280< γ <2400 kg/m3
Jensen (1943) [82] ‫ܧ‬௖ ൌ ͸ ൈ ͳͲ଺ Ȁሺͳ ൅ ʹͲͲͲȀ݂௖ᇱ ሻ psi
14≤ ݂௖ᇱ ≤48 MPa
Pauw (1960) [83] ‫ܧ‬௖ ൌ ͳ͵ǤͺʹߛଵǤ଻ଽ ሺ݂௖ᇱ ሻ଴Ǥସସ
‫ܧ‬௖ ൌ ͵Ǥ͵ʹͲሺ݂௖ᇱ ሻ଴Ǥହ ൅ ͸ǤͻͲͲሻሺ‫ݓ‬ൗʹ͵Ͷ͸ሻ
Carrasquillo, et al.(1981) [75] GPa , kg/m3
‫ܧ‬௖ ൌ ͵Ǥ͵ʹͲሺ݂௖ᇱ ሻ଴Ǥହ ൅ ͸ǤͻͲͲ
Jobse and Mustafa (1984) [79] ‫ܧ‬௖ ൌ ͲǤͳͲ͵ߛଵǤହ ሺඥ݂௖ᇱ ሻ଴Ǥହ
Ahmad and Shah (1985) [78] ‫ܧ‬௖ ൌ ͵Ǥ͵ͺ ൈ ͳͲିହ ൈ ߛ ଶǤହ ሺඥ݂௖ᇱ ሻ଴Ǥ଺ହ
Ravindrarajah et al. (1985)
‫ܧ‬௖ ൌ ͶǤ͸͵Ͳሺ݂௖ᇱ ሻ଴Ǥହ
[116]
Cook (1989) [80] ‫ܧ‬௖ ൌ ͵Ǥʹʹ ൈ ͳͲିହ ൈ ߛ ଶǤହ ሺ݂௖ᇱ ሻ଴Ǥଷଵହ kg/m3 , MPa
Zhang and Gjorv (1991) [81] ‫ܧ‬௖ ൌ ͳǤͳͻሺ݂௖ᇱ ሻଶȀଷ GPa , MPa
Gardner and Zhao(1991) [117] ‫ܧ‬௖ ൌ ͻሺ݂௖ᇱ ሻଵȀଷ ݂௖ᇱ > 27 MPa
Gutierrez and Canovas (1995) య
‫ܧ‬௖ ൌ ͺ͵ͶͲ ඥ݂௖ᇱ
[118]
Rashid et al. (2002) [119] ‫ܧ‬௖ ൌ ͺͻͲͲሺ݂௖ᇱ ሻ଴Ǥଷଷ ʹͲ ൏ ݂௖ᇱ ൏ ͳ͵Ͳ‫ܽܲܯ‬
Levtchitch et el. (2004) [120] ‫ܧ‬௖ ൌ ͳͲͲͲͲሺͳǤͻ ൅ ǤͲͶͷ݂௖ᇱ ሻ ‫ܽܲܯ‬
Leemann and Hoffmann
‫ܧ‬௖ ൌ ͷͶͺͲሺ݂௖ᇱ ሻ଴Ǥହ
(2005) [121]
Kheder and Al-Windawi
‫ܧ‬௖ ൌ ͷǤ͵ʹ͵ሺ݂௖ᇱ ሻ଴Ǥସହଷ GPa, MPa
(2005) [122]
-Best fit with eq. 8-6 in CSA-
ߛ ଵǤହ A23.3(1994)
‫ܧ‬௖ ൌ ൫ ൗʹ͵ͲͲ൯ ሺ͵ͲͲͲሺ݂௖ᇱ ሻ଴Ǥହ ൅ ͸ͻͲͲሻ
-Best fit with eq. 8-6 in CSA-
Soleymani (2006) [76] ߛ ଵǤହ
‫ܧ‬௖ ൌ ൫ ൗʹ͵ͲͲ൯ ሺ͵͵ͲͲሺ݂௖ᇱ ሻ଴Ǥହ ൅ ͸ͻͲͲሻ 23.3(1994)
‫ܧ‬௖ ൌ ͶͷͲͲሺ݂௖ᇱ ሻ଴Ǥହ -worst fit with eq. 8-7 in CSA-
A23.3(1994)
α=55949 Miami oolite limestone,
‫ܧ‬௖ ൌ ߙሺ݂௖ᇱ ሻ଴Ǥହ 62721 for Georgia granite ,
Yanjun Liu (2007) [123]
Ec , ݂௖ᇱ in psi 43777 for stalite lightweight
aggregate
Dinakar (2008) [124] ‫ܧ‬௖ ൌ ͶǤͷͷሺ݂௖ᇱ ሻ଴Ǥହ
Noguchi et al. (2009) [125] ‫ܧ‬௖ ൌ ͵ǤͶͺ ൈ ͳͲିଷ ߛଵǤ଼ଽ ሺ݂௖ᇱ ሻଵȀଷ MPa, kg/m3,40 ≤ ݂௖ᇱ ≤160 MPa
Haranki (2009) [77] ‫ܧ‬௖ ൌ ͵ͳǤͻʹ ൈ ߛଵǤହ ሺ݂௖ᇱ ሻ଴Ǥହ ൅ ͵Ͷͷǡ͵ʹͺ psi , lb/ft3
San Luis Obispo (2011) [126] ‫ܧ‬௖ ൌ ͸Ǥͷͻሺ݂௖ᇱ ሻ଴Ǥଷ଼
Aslani and Nejadi (2012) [127] ‫ܧ‬௖ ൌ Ͷͺ͵ͷሺ݂௖ᇱ ሻ଴Ǥସଽ

93
Chapter 3: Hardened concrete properties

3.5.1.2. Modulus of rupture

Table (3.19) presents number of empirically developed models used to predict the modulus of
rupture of conventional concrete, together with the models in some codes of practice. The
experimental models are the modified form of the models employed in the most frequently used
codes of practice worldwide.

Table 3.19: Existing models to predict the modulus of rupture of conventional concrete

Reference Modulus of rupture (MPa)


EC2-02 [128] ˆ୰ = 0.435(ˆୡᇱ -8)2/3
ˆ୰ ൌ ͲǤ͸ሺˆୡᇱ ሻ଴Ǥହ
CSA A23.3-04 [63]
ˆ୰ ൌ ͲǤ͸ɉሺˆୡᇱ ሻ଴Ǥହ
ACI-363R-92 [129] ˆ୰ ൌ ͲǤͻͶሺˆୡᇱ ሻ଴Ǥହ ʹͳ ൑ ˆୡᇱ ൑ ͺ͵ƒ

Rashid et al. (2002) [119] ˆ୰ ൌ ͲǤͶʹሺˆୡିଵହ଴ ሻ଴Ǥ଺଼ ͷ ൑ ˆୡᇱ ൑ ͳʹͲƒ
IS-456 (BIS, 2000) [130] ˆ୰ ൌ ͲǤ͹ሺˆୡᇱ ሻ଴Ǥହ
Ahmad and shah (1985) [78] ˆ୰ ൌ ͲǤͶͶሺˆୡᇱ ሻ଴Ǥହ ˆୡᇱ < 84 MPa
Jerome (1984) ˆ୰ ൌ ͲǤͶͶሺˆୡᇱ ሻଶȀଷ ˆୡᇱ < 62 MPa
ACI-435 (1968) [131] ˆ୰ ൌ ͲǤ͸ʹ–‘ͲǤͻͻሺˆୡᇱ ሻ଴Ǥହ
Carrasquillo et al. (1981) [132] ˆ୰ ൌ ͲǤͻ͹ሺˆୡᇱ ሻ଴Ǥହ
Bakhsh et al –(1990) [133] ˆ୰ ൌ ͲǤͺሺˆୡᇱ ሻ଴Ǥହ
Raphael, J.M.- (1984) [134] ˆ୰ ൌ ͲǤͶͶሺˆୡᇱ ሻଶȀଷ
Lane-(1998) [135] Ž‘‰ ˆ୰ ൌ ͲǤͷͷͷሺŽ‘‰ˆୡᇱ ሻ െ ͲǤͳͻ͹

The collected 12 models to predict the modulus of rupture of conventional concrete cover the
period since 1968 up to now. The power type relationship is the governing type model.
Limitation on the applied range of compressive strength is effective parameter in some models.

3.5.1.3. Splitting tensile strength

Table (3.20) presents some empirically developed models to predict the splitting tensile
strength of conventional concrete together with the models used in some codes of practice. The
experimental models are the modified form of the models employed in the most frequently used
codes of practice universally.

For prediction of the splitting tensile strength, totally 26 models in the design codes and
empirical equations since 1962 are presented. The power type relationship is used in majority of
the models. Also, the compressive strength is limited in some models.

94
Chapter 3: Hardened concrete properties

3.5.2. Self-compacting Concrete

3.5.2.1. Modulus of elasticity

Self-compacting concrete is a comparatively new construction material in comparison to


conventional and light-weight concrete and the research to discover the most accurate models of
predicting its mechanical properties are still continuing. Therefore, there are fewer models for
modulus of elasticity, splitting tensile strength and modulus of rupture of self-compacting
concrete in literature. Table (3.21) presents some existing models of predicting the modulus of
elasticity of self-compacting concrete in design codes and also the empirical models developed
based on the experimental investigations. The experimental models are the modified form of the
models in the most frequently used codes of practice globally.

Table 3.20: Existing models to predict the splitting tensile strength of conventional concrete

References Splitting tensile strength (MPa)


Akawaza (1953) [136] ˆୡ୲ ൌ ͲǤʹͲͻሺˆୡᇱ ሻ଴Ǥ଻ଷ
Carneriro and Barcellos (1953) [137] ˆୡ୲ ൌ ͲǤͳͺͷሺˆୡᇱ ሻ଴Ǥ଻ଷହ
Vinayaka (1959) [138] ˆୡ୲ ൌ ͲǤͺͺሺˆୡᇱ ሻ଴Ǥ଻ଵ଺
Sen and Desayi (1962) [139] ˆୡ୲ ൌ ͲǤ͸ͺʹሺˆୡᇱ ሻ଴Ǥ଻ଷ
Chapman, G.P., 1968 ) ˆୡ୲ ൌ ͲǤͷͷሺˆୡᇱ ሻ଴Ǥହ
Mirza, S.A et al. (1979) ˆୡ୲ ൌ ͲǤ͵ሺˆୡᇱ ሻଶȀଷ
Carino and Lew (1982) [140] ˆୡ୲ ൌ ͲǤʹ͹ʹሺˆୡᇱ ሻ଴Ǥ଻ଵ
Raphael (1984) [134] ˆୡ୲ ൌ ͲǤ͵ͳ͵ሺˆୡᇱ ሻ଴Ǥ଺଺଻
Ahmad and shah (1985) [78] ˆୡ୲ ൌ ͲǤͶ͸ሺˆୡᇱ ሻ଴Ǥହ Mpa ˆୡᇱ < 84 MPa
Gardner (1990) [141] ˆୡ୲ ൌ ͲǤ͵ͳ͵ሺˆୡᇱ ሻ଴Ǥ଺଺଻
Olokun (1991) [142] ˆୡ୲ ൌ ͲǤʹͲ͸ሺˆୡᇱ ሻ଴Ǥ଺ଽ
଴Ǥହ
ACI-363-R (1992) [129] ˆୡ୲ ൌ ͲǤͷͻˆୡᇱ  ʹͳ ൑ ˆୡᇱ ൑ ͺ͵ƒ
Burg and Ost (1992) [143] ˆୡ୲ ൌ ͹Ǥ͵ሺˆୡᇱ ሻ଴Ǥହ
CEB-FIP (1993) [144] ˆୡ୲ ൌ ͲǤ͵ሺˆୡᇱ ሻଶȀଷ
ACI318 (1995) [60] ˆୡ୲ ൌ ͲǤͷ͸ሺˆୡᇱ ሻ଴Ǥହ
Lane (1998) [135] Ž‘‰ ˆ௖ᇱ ൌ ͳǤͲͻ͸ሺŽ‘‰ˆୡ୲ ሻ ൅ ͲǤͺͷ
EC2-02 (2002) [128] ˆୡ୲ ൌ ͳȀ͵ሺˆୡᇱ െ ͺሻଶȀଷ MPa
CSA A23.3-04 [63] ˆୡ୲ ൌ ͲǤ͸͹ሺˆୡᇱ ሻ଴Ǥହ
NEN-6722 (2000) [145] ˆୡ୲ ൌ ͳ ൅ ͲǤͲͷˆୡᇱ
Rashid et al. (2002) [119] ˆୡ୲ ൌ ͲǤͶ͹ሺˆୡᇱ ሻ଴Ǥହ଺ ͷ ൑ ˆୡᇱ ൑ ͳʹͲƒ

NEN-6722 (2000) [145] ˆୡ୲ ൌ ͳ ൅ ͲǤͲͷˆୡᇱ
ଶ
Hueste et al. (2004) [146] ˆୡ୲ ൌ ͲǤͷͷሺˆୡᇱ ሻ଴Ǥହ
଴Ǥହ
AASHTO (2006) [147] ˆୡ୲ ൌ ͲǤͷͻˆୡᇱ
Dinakar et al. (2008) [124] ˆୡ୲ ൌ ͲǤ͸ͷሺˆୡᇱ ሻ଴Ǥହ
Dinakar (2012) [89] ˆୡ୲ ൌ ͲǤʹͻሺˆୡᇱ ሻ଴Ǥ଺଻ହ
଴Ǥ଻଼
Aslani and Nejadi (2012) [127] ˆୡ୲ ൌ ͲǤͳͻˆୡᇱ

95
Chapter 3: Hardened concrete properties

Table 3.21: Existing models of the modulus of elasticity of self-compacting concrete

References Modulus of elasticity (SI units, Mpa)

‫ܧ‬௖ ൌ ͻͲͲͲ ඥ݂௖ᇱ ݂௖ᇱ > 27 MPa



Gardener and Zhao (1991) [117]
ఊ ଵǤହ
NS 3473 (1992) [148] ‫ܧ‬௖ ൌ ͻǤͷ ቀ ቁ ሺ݂௖ᇱ ሻ଴Ǥଷ , 25 < ݂௖ᇱ < 85 MPa
ଶସ଴଴

ACI 363R-92 [129] ‫ܧ‬௖ ൌ ͵͵ͲͲሺ݂௖ᇱ ሻ଴Ǥହ ൅ ͸ͻͲͲ


య ௙೎ᇲ
CEB-FIP (1993) [144] ‫ܧ‬௖ ൌ ʹͳǤͷ ට GPa, MPa
ଵ଴

‫ܧ‬௖ ൌ ͳͲͲͲͲඥ݂௖ᇱ

EHE (1998) [107]
ACI-318-1999 [60] ‫ܧ‬௖ ൌ Ͷ͵ߛଵǤହ ሺ݂௖ᇱ ሻ଴Ǥହ ൈ ͳͲି଺
Persson (2001) [149] ‫ܧ‬௖ ൌ ͵͹ͷͲሺ݂௖ᇱ ሻ଴Ǥହ
௙ᇲ
EC-2 (2002) [128] ‫ܧ‬௖ ൌ ʹʹሺ ೎ ሻ଴Ǥଷ GPa, MPa
ଵ଴

NBR 6118 (2003) [113] ‫ܧ‬௖ ൌ ͷ͸ͲͲሺ݂௖ᇱ ሻ଴Ǥହ


Huest et al. (2004) [146] ‫ܧ‬௖ ൌ ͷʹ͵Ͳሺ݂௖ᇱ ሻ଴Ǥହ
CSA A23.3-04 [63] ‫ܧ‬௖ ൌ ͶͷͲͲሺ݂௖ᇱ ሻ଴Ǥହ Mpa
Leemann and Hoffmann (2005)[121] ‫ܧ‬௖ ൌ Ͷ͹ͶͲሺ݂௖ᇱ ሻ଴Ǥହ
NZS 3101 (2006) [64] ‫ܧ‬௖ ൌ ͵͵ʹͲሺ݂௖ᇱ ሻ଴Ǥହ ൅ ͸ͻͲͲ MPa
Felekoglu et al.(2007) [150] ‫ܧ‬௖ ൌ ͳǤͷ͹ሺ݂௖ᇱ ሻ଴Ǥ଼ MPa, GPa
ଵǤହ
‫ܧ‬௖ ൌ ͲǤͲͶ͵‫ܭ‬ଵ Ǥ ߛ ሺ݂௖ᇱ ሻ଴Ǥହ MPa and kg/m3 units
AASHTO LRFD (2007)[147]
1440 < γ < 2480 kg/m3 , ݂௖ᇱ <103 MPa
ఊ ଷǤଷ
Atahan et al. (2007) [151] ‫ܧ‬௖ ൌ ͳͲͳͷͲ ቀ ቁ ሺ݂௖ᇱ ሻ଴Ǥଷସଷ MPa , kg/m3
ଶସ଴଴

Dinakar et al. (2008) [124] ‫ܧ‬௖ ൌ ͶͳͺͲሺ݂௖ᇱ ሻ଴Ǥହ


SCC mixtures containing river gravel:
ߛ
‫ܧ‬௖ ൌ ͺͻͺͳሺ݂௖ᇱ ሻ଴Ǥଷ଻଼ ሺ ൗʹͶͲͲሻଵ
‫ܧ‬௖ ൌ ͺʹ͹ͺሺ݂௖ᇱ ሻ଴ଷଽ ൅ ͶʹͶ
For limestone SCC mixture:
ߛ
‫ܧ‬௖ ൌ ͻͶʹͲሺ݂௖ᇱ ሻ଴Ǥଷ଼ଷ ሺ ൗʹͶͲͲሻଶǤଷ
Kim (2008) [152] To compare the MOE between SCC and CC containing limestone
coarse aggregate, the following equation is proposed:
‫ܧ‬௖ ൌ ͹ͳʹͲሺ݂௖ᇱ ሻ଴Ǥସ଴ ൅ ͸ʹͳ
For the unified design equations, the following equation is
proposed for SCC mixtures:
ߛ
‫ܧ‬௖ ൌ ͳͲͳͷ͵ሺ݂௖ᇱ ሻ଴Ǥଷସଷ ሺ ൗʹͶͲͲሻଷǤଷ

Dinakar (2008) [124] ‫ܧ‬௖ ൌ ͶǤͳͺሺ݂௖ᇱ ሻ଴Ǥହ in fsp


ACI 318-08 [153] ‫ܧ‬௖ ൌ Ͷ͹ͲͲሺ݂௖ᇱ ሻ଴Ǥହ MPa
Vilanova et al. (2011) [154] ‫ܧ‬௖ ൌ ͷǤͺͺሺ݂௖ᇱ ሻ଴Ǥସସ GPa, Mpa
Hasan Yıldırım et al.( 2011) [155] ‫ܧ‬௖ ൌ ͷǤͺͷሺ݂௖ᇱ ሻ଴Ǥହ െ ͳ͵Ǥͷ
Debashis Das (2011) [156] ‫ܧ‬௖ ൌ ͵͸ͳͷǤ͸ሺ݂௖ᇱ ሻ଴Ǥହ ൅ ͸ͺͺ͹Ǥ͸ applicable for various ages
Aslani and Nejadi (2012) [127] ‫ܧ‬௖ ൌ ͶͳͷͲሺ݂௖ᇱ ሻ଴Ǥହଶହ

96
Chapter 3: Hardened concrete properties

The collected 24 models to predict the modulus of elasticity of self-compacting concrete are
developed since 1985 up to now. The power type relationship is used in most of the presented
models. Effect of the aggregate type, density of concrete and the compressive strength limit are
some effective parameters included in the model.

3.5.2.2. Modulus of rupture

Table (3.22) presents some existing models of predicting the modulus of rupture of self-
compacting concrete in design codes and also the empirical models developed based on the
experimental investigations. The experimental models are the modified form of the models in
the most frequently used codes of practice worldwide.

Table 3.22: Modulus of rupture of self-compacting concrete in codes of practice and empirical equations

References Modulus of rupture (MPa)


଴Ǥ଺଼
Vilanova et al. (2011) [154] ݂௥ ൌ ͲǤͶ͵݂௖ᇱ
EC-2 (2002) [128] ݂௥ ൌ ͲǤͶ͵ͷሺ݂௖ᇱ െ ͺሻଶȀଷ
CSA A23.3-04 [63] ݂௥ ൌ ͲǤ͸ඥ݂௖ᇱ

NZS-3101 (2006) [64] ݂௥ ൌ ͲǤͺඥ݂௖ᇱ


షఱ ௙ᇲ ሻ
݂௥ ൌ ͵͹ͷ ൈ ͳͲሺଷǤ଻ൈଵ଴ ೎ psi
Atahan et al. (2007) [151]
݂௥ ൌ ʹǤͷͻ ൈ ͳͲ ሺହǤଷൈଵ଴షయ ௙೎ᇲ ሻ

ACI-318-08 [153] ݂௥ ൌ ͲǤͻͶඥ݂௖ᇱ

The models to predict the modulus of rupture of self-compacting concrete are very rare and
just 6 models are found in the published articles and technical reports. These models are
proposed since 2002. Similar to the conventional concrete, the models are of mainly power type
relationship.

3.5.2.3. Splitting tensile strength

Some models to estimate the splitting tensile strength of self-compacting concrete are
presented in Table (3.23). Some models are proposed in codes of practice, while some models
are based on the experimental investigations. The majority of the empirical models are modified
versions of the most frequently used models in the literature.

To predict the splitting tensile strength, totally 21 models developed since 1985 are presented
in the study. Majority of the models are based on the power type relationship between the
compressive strength and modulus of elasticity. In all models, the splitting tensile strength is
derived from the compressive strength.
97
Chapter 3: Hardened concrete properties

3.6. COMPRESSIVE STRESS-STRAIN CURVE IN LIGHTWEIGHT,


CONVENTIONAL AND SELF-COMPACTING CONCRETE

The stress-strain curve is a particularly important graphical presentation of the mechanical


properties of concrete. Experimental stress–strain relationships dealing with different stress
regions are the most usual representation of material properties [157]. The curve is unique for
each type of concrete and is obtained from compression testing of cylinders with a controlled
rate of deformation (displacement) in the concrete specimen subjected to compression. To draw
the stress-strain curve, the deformation (strain) at distinct intervals of compressive loading
(stress) is recorded. These curves expose many of the properties of a material such as the
modulus of elasticity and energy absorption of the concrete specimen. Ductility and brittleness
of the concrete can be found from the stress-strain curves.

Table 3.23: Splitting tensile strength of self-compacting concrete in codes of practice and experimental studies

References Splitting tensile strength (MPa)


଴Ǥହହ
Ahmad and Shah (1985) [78] ݂௖௧ ൌ ͶǤ͵Ͷ݂௖ᇱ
଴Ǥ଺ଽ
Olokun (1991) [142] ݂௖௧ ൌ ͲǤʹͻͷ݂௖ᇱ
଴Ǥହ
Burg and Ost (1992) [143] ݂௖௧ ൌ ͹Ǥ͵݂௖ᇱ
଴Ǥ଺଻
CEB-FIP, (1993) [144] ݂௖௧ ൌ ͲǤ͵Ͳͳ݂௖ᇱ

య ݂௖ᇱ െ ͺ
CEB-FIP (1993) [144] ݂௖௧ ൌ ͳǤͷ͸ ඨቆ ቇ
ͳͲ

య ଶ
EHE (1998) [107] ݂௖௧ ൌ ͲǤʹͳ ට݂௖ᇱ
଴Ǥହ
ACI-318 (1999) [60] ݂௖௧ ൌ ͲǤͷͷ͸݂௖ᇱ

EC-2 (2002) [128] ݂௖௧ ൌ ሺͳൗ͵ ሺ݂௖ᇱ െ ͺሻଶȀଷ

య ଶ
NBR- 6118 (2003) [113] ݂௖௧ ൌ ͲǤ͵ ට݂௖ᇱ
଴Ǥହ
CSA A23.3-04 [63] ݂௖௧ ൌ ͲǤ͸͹݂௖ᇱ
଴Ǥହ
NZS- 3101 (2006) [64] ݂௖௧ ൌ ͲǤͷͶ݂௖ᇱ
଴Ǥ଺ଽ
Atahan et al. (2007) [151] ݂௖௧ ൌ ͲǤʹͻͷ݂௖ᇱ
଴Ǥ଺
Felekoglu et al. (2007) [150] ݂௖௧ ൌ ͲǤͶ͵݂௖ᇱ
଴Ǥହ
Dinakar et al. (2008) [124] ݂௖௧ ൌ ͲǤͺʹ݂௖ᇱ
଴Ǥହ
ACI-318-08 [153] ݂௖௧ ൌ ͲǤͷͻ݂௖ᇱ
Sukumar et al. (2008) [158] ݂௖௧ ൌ ͲǤͲͺͶ͵݂௖ᇱ ൅ ͲǤͺͳͺ
଴Ǥହ
Kim (2008) [152] ݂௖௧ ൌ ͲǤ͸ͺ݂௖ᇱ
Topcu and Uygunog˘lu (2010) [159] ݂௖௧ ൌ ͲǤͲ͸ʹ݂௖ᇱ ൅ ͲǤʹͲͲͻ
଴Ǥ଺଻
Parra et al. (2011) [160] ݂௖௧ ൌ ͲǤʹͺ݂௖ᇱ
଴Ǥ଻ଵ
Vilanova et al. (2011) [154] ݂௖௧ ൌ ͲǤʹ͸݂௖ᇱ
଴Ǥ଼ଽଷ
Aslani and Nejadi (2012) [127] ݂௖௧ ൌ ͲǤͳͳͷ݂௖ᇱ

98
Chapter 3: Hardened concrete properties

Increasing the strain capacity of concrete causes an increase in the area under the stress-strain
diagram and improves the energy absorption capacity. The enhanced energy absorption capacity
changes the concrete into a more ductile material [100]. The greater strain capacity also
improves the resistance to cracking due to restrained shrinkage [161].

Stress–strain curves of various concrete types vary widely; and different compression tests
conducted on the same concrete yield different results, depending upon the temperature of the
specimen and the speed of the loading. Even for the same concrete mixture, these curves are
sensitive to some other parameters. Concrete characteristics and testing conditions are two
major effective parameters in stress–strain curves of concrete. Shape and size of the specimen,
stiffness of the testing machine, type of the strain gauge, gauge length and strain rate are
relevant to testing conditions. Also, the test age and mixture components can change the stress-
strain curve [162].

Mathematical models proposed for the stress-strain relationship of concrete can be divided
into three categories[163]:
a) Continuous function for the ascending and descending branches of the stress-strain curve;
b) Separate curves for the ascending and descending parts meeting at the peak stress;
c) Parabolic curve for the ascending branch together with a bilinear falling curve.

Figure (3.8) shows the general relationship between stress and strain in ductile and brittle
failure of the concrete.

Figure 3.8: Different types of concrete failure in compression

The relationship between the peak compressive stress and corresponding axial strain is
reported as linear in some investigations [163, 164], while some other studies such as Tasnimi

99
Chapter 3: Hardened concrete properties

(2004) [165] propose a polynomial function to describe the relation between compressive stress
and strain in concrete.

Lower stiffness of lightweight aggregate particles and higher cement contents in LWC result
in larger deformations. However, these effects are alleviated by lower density so small scaled
tests in the laboratory can provide pessimistic data, compared to behaviour on site. The stress-
strain relationship for LWC is more linear and brittle than for normal-weight concrete. Such a
response is probably attributable to greater compatibility [100], [166]. Figure (3.9) shows the
general failure mode of LWC and CC. It is evident that for the same grade of concrete, LWC
experiences a brittle failure.

Figure 3.9: Comparing the failure modes of LWC and CC [166]

Yew et al. (2015) compared the stress-strain curves of Oil Palm Shell (OPS) LWC and
conventional concrete. The stress-strain curves of most LWCs for both normal and high-
strength levels were typically linear to the levels approaching 90% or greater of the failure
strength [100].

Lo Monte et al. (2015) [52] investigated the compressive and tensile stress-strain curves of
two mixes of LWC containing EPS beads and one CC mix. They applied thermal cycles at the
temperature of 20°C (room temperature), 150°C, 300°C, 500°C, and 700°C. They found that
both in virgin and slowly heated specimens, at any temperature, the stress-strain curves present
a distinct linear ascending branch, with the strain at the peak of 1.8 to 2.5% at 20°C and 5.5–
8.0% at 700°C, as in CC. The descending branch in conventional concrete was smoother at
higher temperatures.

100
Chapter 3: Hardened concrete properties

Santiago et al. (2009) [167] reported a similar stress-strain curve for all the recycled
aggregate LWC and the natural aggregate concrete. The post-cracking branch of stress-strain
curve of LWC manufactured with waste Ethylene Vinyl Acetate (EVA) was softer. Also, the
slope of the descending branch decreased in their experiment.

3.6.1. Compressive Stress-Strain Curve in Lightweight Concrete

So far, several analytical and numerical models have been developed to simulate the
Compressive Stress-Strain Curve (CSSC) for different types of concrete. Table (3.24) shows
some existing models for calculating CSSC of LWC. Complexity and number of the included
effective parameters affect the precision of the model in calculation of the relation between
stress and strain in these models.

The models to estimate the compressive stress-strain of lightweight concrete are very rare in
the literature. The existing 4 models in the literature since 1975 are collected and compared in
this study. Compressive strength, initial and final slope of the curve, stress and strain at the peak
point and shape factor of the curve are the main parameters in the models. The first developed
models generally contain less parameter and predict the stress-strain behaviour by a simple cure.

101
Chapter 3: Hardened concrete properties

Table 3.24: Existing models of compressive stress-strain curve in lightweight concrete

References CSSC model Notes


ሺ െ  ୮ ሻɂୡ
ɐୡ ൌ ଵȀ୬
൅  ୮ ɂୡ
ሺ െ  ୮ ሻɂୡ ୬
ቈͳ ൅ ൬ ൰ ቉
ɐ଴
Žʹ
ൌെ ɐୡ is the plain concrete stress
ߪ ୮
Žሺ ଵ െ ሻ K : the initial slope of the curve
ߪ଴  െ  ୮
ɂଵ ɂଵ ଶ Kp : the final slope of the curve
Richard and Abbott (1975) ߪଵ ൌ ݂௖ᇱ ቈʹ െ൬ ൰ ቉
ɂ଴ ɂ଴ ɐ଴ : reference stress
[168]
ͲǤ͸ͷɐ଴ n : curve-shape parameter
ɂଵ ൌ
 െ ୮
ɐ଴ ൌ ͳͻǤͳ ൅ ͳǤ͵݂௖ᇱ െ ‫ܭ‬௣ ɂ଴
 ୮ ൌ ͳ͵͹ͶǤͷ െ ͺ͹ͳǤͳ݂௖ᇱ ݂‫݂ݎ݋‬௖ᇱ  ൒ ͳͷ‫ܽܲܯ‬
 ൌ ‫ܧ‬௖ ൌ ͳͺͲǤͻ݂௖ᇱ ൅ ͹͹͹ͲǤ͹
ɂ଴ ൌ ሺͲǤʹ݂௖ᇱ ൅ ͳ͵ǤͲ͸ሻ ൈ ͳͲିସ
‫ ܭ‬ൌ ‫ܧ‬஼ ൌ ͳͺͲǤͻ݂௖ᇱ ൅ ͹͹͹ͲǤ͹
Wang et al. (1978) [169]
Initial slope of curves for Wang et al(1978)
݈݊ʹ
݊ൌെ
݂ଵ ‫ܭ‬௉
݈݊ ൬ െ ൰
݂଴ ‫ ܭ‬െ ‫ܭ‬௉
‫א‬ଵ ‫א‬ଵ ଶ
݂ଵ ൌ ݂௖ᇱ ቈʹ െ൬ ൰ ቉
‫א‬଴ ‫א‬଴
ɐ଴
‫א‬ଵ ൌ ݂‫ܥܥݎ݋‬
‫ ܭ‬െ ‫ܭ‬௉
ͲǤ͸ͷ݂଴
Almussalam and alsayed ‫א‬ଵ ൌ ݂‫ܥܹܮݎ݋‬
‫ ܭ‬െ ‫ܭ‬௉
(1995) [163]
ɐ଴ ൌ ͷǤ͸ ൅ ͳǤͲʹ݂௖ᇱ െ ‫ܭ‬௉ ߳௖ ݂‫ܥܥݎ݋‬
ɐ଴ ൌ ͳͻǤͳ ൅ ͳǤ͵݂௖ᇱ െ ‫ܭ‬௉ ߳௖ ݂‫ܥܹܮݎ݋‬
‫ܭ‬௉ ൌ ͷͶ͹Ͳ െ ͵͹ͷ݂௖ᇱ ݂‫݂ݎ݋‬௖ᇱ ൑ ͷͷ‫ܽܲܯ‬
‫ܭ‬௉ ൌ ͳ͸͵ͻͺǤʹ͵ െ ͸͹͸Ǥͺʹ݂௖ᇱ 
݂‫݂ݎ݋‬௖ᇱ ൐ ͷͷ‫ܽܲܯ‬
‫ܭ‬௉ ൌ ͳ͵͹ͶǤͷ െ ͺ͹ͳǤͳ݂௖ᇱ 
݂‫݂ݎ݋‬௖ᇱ ൒ ͳͷ‫ܥܹܮݎ݋݂ܽܲܯ‬
ஒሺகȀக ሻ
ˆୡᇱ బ
ߪ௖ ൌ ݂݅ɂ ൑  ɂ଴  For permeable concrete
ሾሺȾ െ ͳ ൅ ሺɂȀɂ଴ ሻஒ ሿ
ɐୡ ǡ ɂǣ …‘…”‡–‡•–”‡•• and strain
ˆୡᇱ Ǥ ȾሺɂȀɂ଴ ሻ
ߪ௖ ൌ ݂݅ɂ ൐  ɂ଴  ˆୡᇱ ǡ ɂ଴ ǣ Maximum stress and
ሾሺȾ െ ͳ ൅ ሺɂȀɂ଴ ሻ୬ஒ ሿ
ͳ corresponding strain
Hussin et al. (2013) [170] Ⱦൌ
݂௖ᇱ β: Material parameter
ͳെ
‫ܧ‬଴ ߝ଴
‫ܧ‬଴ : Initial tangent modulus of
݊ ൌ ͲǤͲͲͳʹܲଶ െ ͲǤͲͺͶ͸ܲ ൅ ͳǤͷͷͳ
elasticity
ˆୡᇱ ൌ ͳͶͻǤͷͷ݁ ି଴Ǥଵ௉
P: porosity of concrete
ɂ଴ ൌ ሺʹ݂௖ᇱ ሻ ൈ ͳͲିହ ൅ ͲǤͲͲͲͺ

102
Chapter 3: Hardened concrete properties

3.6.2. Compressive Stress-Strain Models in Conventional Concrete

Table (3.25) shows the existing models of predicting the relationship between stress and
strain under compression in conventional concrete.

Table 3.25: Existing models of compressive stress-strain curve in conventional concrete

References CSSC model Notes


ɂ ɂ
ߪ௖ ൌ ˆୡᇱ ൤ʹ ൬ ൰ െ ሺ ሻଶ ൨Ǣ ɂ ൑ ɂ଴
ɂ଴ ɂ଴
ɂ െ ɂ଴
Hognestad (1951) [171] ߪ௖ ൌ ˆୡᇱ ൤ͳ െ ͲǤͳͷ ൬ ൰൨ ǡɂ ˆୡᇱ ǡ ɂ଴ ǣ Maximum stress and strain
ɂ୳ െ ɂ଴
൐ ɂ଴
‫ܧ‬௖ ൌ ͳʹ͸ͺͲ ൅ Ͷ͸Ͳ݂௖ᇱ 

Smith and Young (1955) [172] ሺଵିக ሻ
ߪ௖ ൌ ‫ܧ‬௖ ɂɂ଴ బ

‫ܧ‬௖ ɂ
Desayi and Krishnan (1964) [173] ߪ௖ ൌ ൦ ൪
ɂ ଶ
ͳ൅ቀ ቁ
ɂ଴

‫ܧ‬௖ ɂ
ߪ௖ ൌ ˆୡᇱ ൦ ൪
‫ܧ‬௖ ɂ ɂ ଶ
Saenz (1964) [174] ͳ ൅ ൬ െ ʹ൰ ቀ ቁ ൅ ቀ ቁ
‫ܧ‬௣ ɂ଴ ɂ଴
ˆୡᇱ
‫ܧ‬௣ ൌ
ɂ଴
ሺ െ  ୮ ሻɂୡ
ɐୡ ൌ ଵȀ୬
൅  ୮ ɂୡ
ሺ െ  ୮ ሻɂୡ ୬
ቈͳ ൅ ൬ ɐ଴
൰ ቉

Žʹ
ൌെ
ߪଵ ୮
Žሺ െ ሻ ɐୡ is the plain concrete stress
ߪ଴  െ  ୮
ɂଵ ɂଵ ଶ K : the initial slope of the curve
ߪଵ ൌ ݂௖ᇱ ൤ʹ െ൬ ൰ ൨
ɂ଴ ɂ଴ Kp : the final slope of the curve
Richard and Abbott (1975) [168] ɐ଴ ɐ଴ : reference stress
ɂଵ ൌ
 െ ୮
n : curve-shape parameter
ɐ଴ ൌ ͷǤ͸ ൅ ͳǤͲʹ݂௖ᇱ െ ‫ܭ‬௣ ɂ଴
(for normal and high strength concrete)
 ୮ ൌ ͷͶ͹Ͳ െ ͵͹ͷ݂௖ᇱ Ǣ݂௖ᇱ  ൑ ͷͷ‫ܽܲܯ‬
 ୮ ൌ ͳ͸͵ͻͺǤʹ͵ െ ͸͹͸Ǥͺʹ݂௖ᇱ 
݂௖ᇱ  ൐ ͷͷ‫ܲܯ‬
 ൌ ‫ܧ‬௖ ൌ ͵͵ʹͲඥ݂௖ᇱ ൅ ͸ͻͲͲ
ɂ଴ ൌ ሺͲǤʹ݂௖ᇱ ൅ ͳ͵ǤͲ͸ሻ ൈ ͳͲିସ
ˆୡᇱ Ǥ ȾሺɂȀɂ଴ ሻ ɐୡ ǡ ɂǣ …‘…”‡–‡•–”‡•• and strain
ߪ௖ ൌ
ሾሺȾ െ ͳሺɂȀɂ଴ ሻஒ ሿ
ˆୡᇱ ǡ ɂ଴ ǣ Maximum stress and corresponding
Carreira and Chu (1985) [164] ͳ
Ⱦൌ strain
ˆୡᇱ
൤ͳ െ ሺ ሻ൨ β: Material parameter
‫ܧ‬଴ ɂ଴
ɂ଴ ൌ ሺ͹Ǥͳ݂௖ᇱ ൅ ͳ͸ͺͲሻ ൈ ͳͲି଺ ‫ܧ‬଴ : Initial tangent modulus of elasticity
ߪ௖ ൌ ‫ܧ‬௖ ɂǢ ɂ ൑ ɂ଴ 
க ିக
Mazars and Pijaudier- Cabot (1989) ሺ బᇲ ሻ
ߪ௖ ൌ ቈሺͳ െ ߙሻɂ଴ ൅ ߙɂ‡ ఌ೎ ቉ ‫ ܧ‬Ǣ ɂ
௖ ൐ ɂ଴
[175]
ˆୡᇱ െ ‫ܧ‬௖ ɂ଴
Ƚൌ கబ
ᇲ ିଵ
‫ܧ‬௖ ሺߝ௖ᇱ ݁ ఌ೎ െ ɂ଴ ሻ

103
Chapter 3: Hardened concrete properties

Table 3.25: Existing models of compressive stress-strain curve in conventional concrete (continued)

References CSSC model Notes


ɂ
ሺ ሻ
ᇱ ɂ଴
ߪ௖ ൌ ˆୡ ቎ ɂ ቏
 െ ͳ ൅ ሺ ሻ௡௞
Collins and Mitchell ɂ଴
(1991) [176] ɂ ൑ ɂ଴ Ǣ ݇ ൌ ͳ
ˆୡᇱ
ɂ ൐ ɂ଴ Ǣ ݇ ൌ ͲǤ͸͹ ൅ ǡ ‫ܽܲܯ‬
͸ʹ
ˆୡᇱ ˆୡᇱ ݊
݊ ൌ ͲǤͺ ൅ ǡ ɂ଴ ൌ  
ͳ͹ ‫ܧ‬௖ ݊ െ ͳ
‫ܧ‬ ɂ ɂ ଶ
ቀ ௜௧ ቁ ቀ ቁ െ ቀ ቁ
‫ܧ‬ ɂ଴ ɂ଴
ߪ௖ ൌ ˆୡᇱ ൦ ଴ ൪Ǣ Ͳ ൑ ɂ ൑ ɂ୫ୟ୶
‫ܧ‬௜௧ ɂ
ͳ ൅ ቀ െ ʹቁ ቀ ቁ
‫ܧ‬଴ ɂ଴

ͳ ‫ܧ‬௜௧ ͳ ͳ ‫ܧ‬௜௧ ͳ
ɂ୫ୟ୶ ൌ ɂ଴ ቎ ൬ ൅ ͳ൰ ൅ ඨ ൬ ൅ ͳ൰ െ ቏
ʹ ‫ܧ‬଴ ʹ ʹ ‫ܧ‬଴ ʹ
ˆୡᇱ
ߪ௖ ൌ Ǣ ɂ
ͳ ʹ ɂ ଶ Ͷ ɂ
൤ ߮െ ൨ቀ ቁ ൅ ൤ െ ߮൨
CEB-FIP (1990) [103] ሺɂ୫ୟ୶ Ȁɂ଴ ሻ ሺɂ୫ୟ୶ Ȁɂ଴ ሻ ɂ଴
ଶ ሺɂ୫ୟ୶ Ȁɂ଴ ሻ ɂ଴
൐ ɂ୫ୟ୶

ɂ୫ୟ୶ ଶ ‫ܧ‬௜௧ ɂ ‫ܧ‬


Ͷ ൤ቀ ቁ ቀ െ ʹቁ ൅ ʹ ቀ ୫ୟ୶ ቁ െ ௜௧ ൨
ɂ଴ ‫ܧ‬଴ ɂ଴ ‫ܧ‬଴
߮ൌ ଶ
ɂ୫ୟ୶ ‫ܧ‬௜௧
ቂቀ ቁ ቀ െ ʹቁ ൅ ͳቃ
ɂ଴ ‫ܧ‬଴
య ᇱ ˆୡᇱ
‫ܧ‬௜௧ ൌ ʹͳͷͲͲඥˆୡ ȀͳͲǡ‫ܧ‬଴ ൌ ǡ ɂ଴ ൌ ͲǤͲͲʹʹ
ͲǤͲͲʹʹ

ɐୡ ǡ ɂǣ …‘…”‡–‡•–”‡•• and
ˆୡᇱ Ǥ ȾሺɂȀɂ଴ ሻ
ߪ௖ ൌ strain
ሾሺȾ െ ͳ ൅ ሺɂȀɂ଴ ሻஒ ሿ
ˆୡᇱ ǡ ɂ଴ ǣ Maximum stress and
(Almusallam & Alsayed ͳ
Ⱦൌ corresponding strain
(1995) [163] ˆᇱ
൤ͳ െ ሺ ୡ ሻ൨ Β: Material parameter
‫ܧ‬଴ ɂ଴
଴Ǥସସ ‫ܧ‬଴ : Initial tangent modulus
ɂ଴ ൌ ሺͲǤ͵ͻͺ݂௖ᇱ െ ͸Ǥ͹Ͷͺሻ ൈ ͳͲିହ
of elasticity
݇ଵ ȾሺɂȀɂ଴ ሻ For normal and high
ߪ௖ ൌ ˆୡᇱ ൤ ൨
݇ଵ Ⱦ െ ͳ ൅ ሺɂȀɂ଴ ሻ௞మஒ strength concrete
݇ଵ ൌ ሺͷͲȀˆୡᇱ ሻଷ , ݇ଶ ൌ ሺͷͲȀˆୡᇱ ሻଵǤଷ ; ˆୡᇱ ൑ ͷͲ‫ܽܲܯ‬ ɐୡ ǡ ɂǣ …‘…”‡–‡•–”‡•• and
݇ଵ ൌ ݇ଶ ൌ ͳ , for ˆୡᇱ ൐ ͷͲ‫ܽܲܯ‬ strain
Wee et al. (1996) [177] ɂ଴ ൌ ͲǤͲͲͲ͹ͺሺ݂௖ᇱ ሻ଴Ǥଶହ ˆୡᇱ ǡ ɂ଴ ǣ Maximum stress and
ͳ corresponding strain
Ⱦൌ
ˆᇱ β: Material parameter
൤ͳ െ ሺ ୡ ሻ൨
‫ܧ‬଴ ɂ଴ ‫ܧ‬଴ : Initial tangent modulus

‫ܧ‬௖ ൌ ͳͲʹͲͲඥ݂ ᇱ
௖ of elasticity
ɐୡ ǡ ɂǣ …‘…”‡–‡•–”‡•• and
ˆୡᇱ Ǥ ȾሺɂȀɂ଴ ሻ
ߪ௖ ൌ strain
ሾሺȾ െ ͳሺɂȀɂ଴ ሻஒ ሿ
ˆୡᇱ ǡ ɂ଴ ǣ Maximum stress and
ͳ
Tasnimi(2004) [162] Ⱦൌ corresponding strain
ˆᇱ
൤ͳ െ ሺ ୡ ሻ൨ β: Material parameter
‫ܧ‬଴ ɂ଴
଴Ǥସସ ‫ܧ‬଴ : Initial tangent modulus
ɂ଴ ൌ ሺ͸ͷǤͷ͹݂௖ᇱ െ ͸Ǥ͹Ͷͺሻ ൈ ͳͲିହ
of elasticity
ɂ

Ƚቀ
ɂ଴
ߪ௖ ൌ ˆୡᇱ ቎ ɂ ቏
ͳ ൅ Ⱦቀ ቁ
ɂ଴
Chandrasekhar et al. க
ൌ ͳ‫ߪ ݄݊݁ݐ‬௖ ൌ ˆୡᇱ
கబ
(2011) [178] க க
ൌ ͳǢ ݀(ߪ௖ Ȁˆୡᇱ ሻሺ ሻ ൌ ͳ
கబ கబ
Ascending portion: α=1.486 , β=0.486
Descending portion: α=2.462 , β=1.462

104
Chapter 3: Hardened concrete properties

Total number of 13 models to estimate the compressive stress-strain curve in the


conventional concrete is collected. There are major differences in the simplicity and complexity
of the models to include different parameters in the model. The oldest model is developed in
1951. Wide range of the limitation of the applicable compressive strength, boundary of the
linear part of the curve, peak stress and strain and the ultimate strain are the main parameters in
almost all the models. The first developed models generally contain less parameter and predict
the stress-strain behaviour by a simple curve with limited number of parameters.

3.6.3. Compressive Stress-Strain Models in Self-compacting Concrete

Table (3.26) shows the existing models of predicting the relationship between stress and
strain under compression in self-compacting concrete.
Compared to the conventional and lightweight concrete, self-compacting concrete is a new
construction material and the models to estimate the compressive-stress-strain are very rare in
the literature. In this study a total number of 4 models are presented to estimate the variation of
stress-strain diagram in self-compacting concrete under compression. The included parameters
are similar to the conventional concrete. However, the applied coefficients are different. The
stress and strain at the peak point, initial slope, the ultimate stain and the shape factor of curve
are the main parameters in the developed models.

105
Chapter 3: Hardened concrete properties

Table 3.26: Existing models of compressive stress-strain curve for self-compacting concrete
References CSSC model

ʹǤͺ͸͸ʹሺɂȀɂ଴ ሻ
ߪ௖ ൌ ˆୡᇱ ቎ ɂ ቏  ‫ ׷‬ɂ ൑ ɂ଴
ͳ ൅ ͲǤͺ͸͸ʹ ቀ ቁ ൅ ሺɂȀɂ଴ ሻଶ
ɂ଴
Prasad et al. (2009 ) [179]
ʹǤͺ͸͸ʹሺɂȀɂ଴ ሻ
ߪ௖ ൌ ˆୡᇱ ቎ ɂ ቏  ‫ ׷‬ɂ ൐ ɂ଴
ͳ ൅ ͲǤͺ͸͸ʹ ቀ ቁ ൅ ሺɂȀɂ଴ ሻଶ
ɂ଴
ɂ ɂ
Ƚ ቀ ቁ ൅ ሺȾ െ ͳሻሺ ሻଶ
ᇱ ɂ଴ ɂ଴
ߪ௖ ൌ ˆୡ ቎ ɂ ɂ ቏
ͳ ൅ ሺȽ െ ʹሻ ቀ ቁ ൅ Ⱦሺ ሻଶ
ɂ଴ ɂ଴
ɂ ɂ
ሺ ଼଴ ሻଶ െ ሺͲǤʹߙ ൅ ͳǤ͸ሻሺ ଼଴ ሻ
ɂ଴ ɂ଴
Ⱦൌ ɂ଼଴ ଶ
Kumar et al. (2011) [180] ͲǤʹሺ ሻ
ɂ଴
ɂ଴
Ƚ ൌ ‫ܧ‬଴ ᇱ Ǣ ݇ ൌ ͳǤͺͷ െ ͲǤͲʹͷˆୡᇱ
ˆୡ

ɂ଴ ൌ ͲǤͲͲͲ͸ ඥ݂௖ᇱ Ǣ‫ܧ‬௖ ൌ ͷ͵ͲͲሺ݂௖ᇱ ሻ଴Ǥସ଺ 
ɂ଼଴ ൌ ሺͲǤͲͲͳͷሻሺ݂௖ᇱ ሻ଴Ǥଵ଺ 
ɂ
Ƚቀ ቁ
ɂ଴
ߪ௖ ൌ ˆୡᇱ ቎ ɂ ቏
ͳ ൅ Ⱦቀ ቁ
ɂ଴

ൌ ͳ‫ߪ ݄݊݁ݐ‬௖ ൌ ˆୡᇱ
Chandrasekhar et al. (2011) [178] கబ
க க
ൌ ͳǢ ݀(ߪ௖ Ȁˆୡᇱ ሻሺ ሻ ൌ ͳ
கబ கబ

Ascending portion: α=1.331 , β=0.331


Descending portion: α=2.462 , β=1.462
ߝ
݊ ൬ ᇱ൰
ߝ௖
ߪ௖ ൌ ݂௖ᇱ
ߝ ௡
݊ െ ͳ ൅ ൬ ᇱ൰
ߝ௖
‫ܧ‬௦௘௖ ି଴Ǥ଻ସ
݊ ൌ ݊ଵ ൤ͳǤͲʹ െ ͳǤͳ͹ ൬ ൰൨ Ǣ ߝ ൑ ߝ௖ᇱ
‫ܧ‬௖
݊ ൌ ݊ଶ ൌ ݊ଵ ൅ ሺߣ ൅ ʹͺߤሻǢ ߝ ൐ ߝ௖ᇱ 
Aslani and Nejadi (2012) [127] ߣ ൌ ሺͳ͵ͷǤͳ͸ െ ͲǤͳ͹ͶͶ݂௖ᇱ ሻି଴Ǥସ଺
ͻͳͳ
ߤ ൌ ͲǤͺ͵‡š’ሺെ ሻ
݂௖ᇱ
݂ᇱ
‫ܧ‬௦௘௖ ൌ ௖൘ߝ ᇱ

݂௖ᇱ ߰ ݂௖ᇱ
ߝ௖ᇱ ൌ ቆ ቇ൬ ൰ǡ߰ ൌ ൅ ͲǤͺ
‫ܧ‬௖ ߰ െ ͳ ͳ͹

106
Chapter 3: Hardened concrete properties

3.7. STEEL-CONCRETE BOND CHARACHTERISTICS

The bond stress between reinforcing bars and the surrounding concrete is the shearing force
that results from the chemical adhesion, friction and mechanical interaction between concrete
and steel. The first two parameters are typical in plain and deformed bars, and the last parameter
is mostly the matter of deformed bars.

3.7.1. Bond Characteristics in Lightweight Concrete

The bond strength of LWC is similar to that of normal-weight concrete. Anchorage in LWC
is lower than that of conventional concrete due to the lower bearing strength caused by the
relative weakness of lightweight aggregates [166].

According to BS-8110 [72], the bond stresses due to determined lap and anchorage lengths
for LWC can be taken as being as conservative as 80% of those for the same grade of
compressive strength in CC. Table (3.27) shows some existing models of predicting the bond
between reinforcing steel bars and surrounding concrete in lightweight concrete.

Despite numerous models to predict different mechanical properties of concrete, only one
model is found in the literature to describe the relationship between bond stress between
reinforcement and concrete. This model is particularly developed for lightweight concrete in
2013. Bar diameter, compressive strength of concrete and the embedded length of bar are the
main included parameters in the models.

Table 3.27: Models to estimate the steel-concrete bond stress in light-weight concrete

Reference Bond stress (MPa) Notes


͵͹Ǥͷ db: rebar diameter
߬௠௔௫ ൌ ൤ െ ͻǤͶ൨ ඥ݂௖ᇱ
Kim et al. (2013) [181] ሺ݀௕ ൅ ݈ௗ ሻ଴Ǥଶହ
ld: embedded length

3.7.2. Bond Characteristics in Conventional Concrete

Table (3.28) shows some existing models of predicting the bond stress between reinforcing
steel bars and surrounding concrete in conventional concrete.

The bond stress between the concrete and embedded steel bar is widely investigated in the
literature since 1977 and 25 models are collected in this study. Diameter and embedded length
of steel bar, compressive strength of surrounding concrete and the concrete cover are almost the
main parameters to define the bond stress in the models. However, in some models the tensile
strength of steel bar, plane or deformed bar and the ribs on the bar are included.
107
Chapter 3: Hardened concrete properties

Table 3.28: Existing models to estimate the steel-concrete bond stress in conventional concrete

References Bond model Notes


db: rebar diameter
ܿ ݀௕
߬௠௔௫ ൌ ൤ͳǤʹʹ ൅ ͵Ǥʹ͵ ൅ ͷ͵ ൨ ሺ݂௖ᇱ ሻ଴Ǥହ ld: embedded length
Orangun et al (1977) ݀௕ ݈ௗ
cm: smaller of clear bottom cover to
[182] ‫ܥ‬௠ ݀௕
߬௠௔௫ ൌ ൤ͲǤͳ ൅ ͲǤʹͷ ൬ ൰ ൅ ͶǤͳͷሺ ሻ൨ ሺ݂௖ᇱ ሻ଴Ǥହ main reinforcement or half clear
݀௕ ݈ௗ
spacing between bars
Kemp and Wilhelm ܿ ‫ܣ‬௧ ݂௬
߬௠௔௫ ൌ ൬ͲǤͷͷ ൅ ͲǤʹͶ ൰ ሺ݂ ᇱ ሻ଴Ǥହ ൅ ͲǤͳͻͳ
(1979) [183] ݀௕ ௖ ‫ݏ‬Ǥ ݀௕

Eligehausen et al. ܿǤ ݂௖ᇱ


߬௠௔௫ ൌ ͲǤ͹ͷඨ
(1983) [184] ݀௕
ܿ
Kemp (1986) [185] ߬௠௔௫ ൌ ʹ͵ʹʹ ൅ ʹǤ͹ͳ͸ ሺ݂ ᇱ ሻ଴Ǥହ
݀௕ ௖
Robins and Austin
߬௠௔௫ ൌ ͳǤ͵ͺඥ݂௖ᇱ
(1986) [186]
Chapman and Shah ܿ ݀௕
߬௠௔௫ ൌ ൤͵Ǥͷ ൅ ͵ǤͶ ൅ ͷ͹ ൨ ሺ݂௖ᇱ ሻ଴Ǥହ
(1987) [187] ݀௕ ݈ௗ
ܿ ݀௕
Harajli (1994) [188] ߬௠௔௫ ൌ ൤ͳǤʹ ൅ ͵ ൅ ͷͲ ൨ ሺ݂௖ᇱ ሻ଴Ǥହ
݀௕ ݈ௗ
Huang et al. (1996)
߬௠௔௫ ൌ ͲǤͶͷ݂௖ᇱ
[189]
ܿ
Esfahani and ൅ ͲǤͷ
݀௕
߬௠௔௫ ൌ ቌʹǤ͸ͻͷ ൈ ܿ ቍ ሺ݂௖ᇱ ሻ଴Ǥହ
Rangan (1998) [190] ൅ ͷǤͷ
݀௕
ܿ ݀௕
߬௠௔௫ ൌ ቈͲǤͳ ൅ ͲǤʹͷ ൅ ͶǤʹ
Pillai et al. (1999) ݀௕ ݈ௗ
[191] ‫ܣ‬௧௥ ݂௬
൅ ͲǤͲʹͶ ቉ ሺ݂௖ᇱ ሻ଴Ǥହ
‫ݏ‬Ǥ ݀௕

CEB-FIP (1990) ߬௠௔௫ ൌ ʹǤͷඥ݂௖ᇱ Confined


[103] ߬௠௔௫ ൌ ʹǤͲඥ݂௖ᇱ Unconfined

Tmax: ultimate bond force (N)


ܶ௠௔௫ As: cross section area
ൌ ሾͳǤͶͶ݈ௗ ሺܿ ൅ ͲǤͷ݀௕ ሻ ൅ ͷ͸Ǥ͵‫ܣ‬௦ ሿ
Zuo and Darwin ሺ݂௖ᇱ ሻ଴Ǥଶହ
Cm: maximum of clear bottom
(2000) [192] ‫ܥ‬௠
ൈ ൬ͲǤͳ ൅ ͲǤͻ൰ cover to main reinforcement and
ܿ
half clear spacing between bars
ACI-408R (2003) ඥ݂௖ᇱ
߬௠௔௫ ൌ ʹͲǤʹ͵
[193] ݀௕
Sancak (2009)
߬௠௔௫ ൌ ͲǤ͸͵ሺ݂௖ᇱ ሻ଴Ǥ଻ଵସ At 90 days
[194]
Tanyildizi (2009)
߬௠௔௫ ൌ ͲǤͳ͹ͻ͸݂௖ᇱ െ ͲǤ͵ͺͶͷ
[195]
Desnerck (2010) ܿ
߬௠௔௫ ൌ ൬ͳǤͻͶ ൅ ͲǤͲǤʹͻͳ ൰ ඥ݂௖ᇱ
[196] ݀௕

108
Chapter 3: Hardened concrete properties

Table 3.28: Models to estimate the steel-concrete bond stress in conventional concrete (continued)

References Bond model Notes


ܿ ଴Ǥ଺ ݀௕
߬௠௔௫ ൌ ቈͲǤ͹ ൬ ൰ ൅ Ͷ ൬ ൰቉ ሺ݂௖ᇱ ሻ଴Ǥଶଷ
Aslani and Nejadi ݀௕ ݈ௗ Plain rebar
(2012) [197] ܿ ଴Ǥ଺ ݀௕ Deformed rebar
߬௠௔௫ ൌ ቈͲǤ͹ ൬ ൰ ൅ Ͷ ൬ ൰቉ ሺ݂௖ᇱ ሻ଴Ǥଶଷ
݀௕ ݈ௗ
Kim et al. (2013) ͵͹Ǥͷ
߬௠௔௫ ൌ ൤ െ ͻǤͶ൨ ඥ݂௖ᇱ
[181] ሺ݀௕ ൅ ݈ௗ ሻ଴Ǥଶହ

Nilson (1971) [198] ߬௠௔௫ ൌ ሺͳǤͶ͵‫ ݔ‬൅ ͳǤͷሻඥ݂௖ᇱ


ACI-408R (2003) ඥ݂௖ᇱ db :bar diameter of the (mm)
߬௠௔௫ ൌ ʹͲǤʹ͵
[193] ݀௕ f’c : compressive strength (MPa).

Kco =c/db combined parameter of


Wu and Zhoa (2012) ʹǤͷ bar diameter and concrete cover
߬௠௔௫ ൌ ൤ ൨ ඥ݂௖ᇱ
[199] ͳ ൅ ͵Ǥͳ݁ ି଴Ǥସ଻ሺ௄೎೚ାଷଷ௄ೞ೟ሻ
Kst =Ast/nSst db confinement
factor
߬௠௔௫ ൌ ሾͲǤͳͺͶ͸ͺ ൅ ͲǤͲͻʹ͵Ͷሺܿ Τ݀௕ ሻ
൅ ͳǤ͸Ͳ͵ͺሺ݀௕ Τ݈ௗ ሻ C: min(cx, cy, cs/2)
൅ ͲǤ͸͵ͳͺሺ݄௥ Τ‫ݏ‬௥ ሻሿඥ݂௖ᇱ ݂‫݂ݎ݋‬௖ᇱ
db: bar diameter
Diab et al. ൏ ͺͲ‫ܽܲܯ‬
ld: embedded length
(2014)[200] ߬௠௔௫ ൌ ሾͲǤͲͺʹ͸ʹ ൅ ͲǤͲͻʹ͵Ͷሺܿ Τ݀௕ ሻ
൅ ͳǤ͸Ͳ͵ͺሺ݀௕ Τ݈ௗ ሻ hr: rib height
൅ ͲǤ͸͵ͳͺሺ݄௥ Τ‫ݏ‬௥ ሻሿඥ݂௖ᇱ ݂‫݂ݎ݋‬௖ᇱ sr: rib spacing
൒ ͺͲ‫ܽܲܯ‬

BS-8110 (1997) [72] ߬௠௔௫ ൌ ͲǤͶඥ݂௖ᇱ

߬௠௔௫ ൌ ʹǤʹͷߟଵ ߟଶ ݂௖௧


ଶൗ
ͲǤʹ݂௖ᇱ ଷ
݂௖௧ ൌ Ǣ݂௖ᇱ ൑ ͸Ͳ‫ܽܲܯ‬
EN-92 (2002) [128] ߛܿ
ሺͳ ൅ ሺ݂௖ᇱ ൅ ͺሻ
ͳǤͶͺ Ž ൬ ൰
ͳͲ
݂௖௧ ൌ Ǣ݂௖ᇱ ൑ ͸Ͳ‫ܽܲܯ‬
ߛܿ
ଶȀଷ
Marti et al. (1998) ߬௠௔௫ ൌ ͲǤ͸݂௖ᇱ  ‫ߝ  ׷‬௦ ൑ ߝ௬
ଶȀଷ Tension chord model
[201] ߬௠௔௫ ൌ ͲǤ͵݂௖ᇱ  ‫ߝ  ׷‬௦ ൐ ߝ௬
Oh and Kim (2007) ଴Ǥ଺
߬௠௔௫ ൌ ʹǤͷ݂௖ᇱ
[202]

3.7.3. Bond Characteristics in Self-compacting Concrete

Table (3.29) shows some existing models of predicting the steel-concrete bond behaviour in
conventional concrete. Only 3 empirical models since 2007 are found in the literature to
describe the relationship between the bond stress between the steel bars and the surrounding

109
Chapter 3: Hardened concrete properties

self-compacting concrete. Bar diameter, compressive strength of concrete, and the embedded
bar length are the effective parameters in the models.

Table 3.29: Models to estimate the steel-concrete bond stress in self-compacting concrete

References Bond model Notes


ܿ ஻
߬௠௔௫ ൌ ‫ ܣ‬൬ ൰ ൈ ሺ݂௖ᇱ ሻఈ
݀௕
ߙ ൌ ͲǤʹͳǡ ͲǤͷͺǡ ͲǤͶͷ
Lightweight SCC Normal
Plain, deformed and GFRP
SCC
Bae (2006) [203] rebar
deformed GFRP plain deformed
GFRP: Glass Fibre
A 0.85 0.48 0.3 0.74
Reinforced Polymer
B 0.17 0.68 0.88 0.52
Embedded length < 15 db
ܿ
Desnerck (2010) [196] ߬௠௔௫ ൌ ൬ͳǤ͹͸ʹ ൅ ͲǤͷͳͶ ൰ ඥ݂௖ᇱ
݀௕
ܿ ଴Ǥ଺ ݀௕
߬௠௔௫ ൌ ቈͲǤ͸͹ͻ ൬ ൰ ൅ ͵Ǥͺͺ ൬ ൰቉ ሺ݂௖ᇱ ሻ଴Ǥହହ
Aslani and Nejadi ݀௕ ݈ௗ Plain rebar
(2012) [197] ܿ ଴Ǥ଺ ݀௕ Deformed rebar
߬௠௔௫ ൌ ቈͲǤ͸͹ʹ ൬ ൰ ൅ ͶǤͺ ൬ ൰቉ ሺ݂௖ᇱ ሻ଴Ǥହହ
݀௕ ݈ௗ

3.8. SUMMARY

The lightweight concrete containing Expanded Poly-Styrene (EPS) beads is the particular
type of concrete that used to manufacture the testing samples and slab specimens of this study.
General specifications of the EPS beads are given in the technical documentations. Almost all
the existing investigations about EPS lightweight concrete since 1976, including the
experimental data are presented. The collected data contains the information of curing method,
type of fine and coarse aggregates, mineral fillers, chemical admixtures, and fibres in each
experiment. The mixture proportions including the size and volume of EPS beads in all
experiments, together with the obtained density and compressive strength of the concrete, are
presented.

Mechanical properties of EPS lightweight concrete from 154 mixture design in 55


experimental programs are also described. In the presented tables, the volume of EPS beads
corresponding to the reported mechanical properties is also given.

The lightweight concrete is classified into structural and non-structural types in many codes
of practice. Therefore, the collected data was investigated in the corresponding categories. From

110
Chapter 3: Hardened concrete properties

the regression analysis of the experimental data, new relationships are developed between the
compressive strength and other mechanical properties of EPS lightweight concrete.

In the absence of reliable testing facility, mechanical properties are extracted from the
existing models in the literature. Although there are a wide range of relationships in the codes of
practice and empirical equations for predicting the mechanical properties of lightweight and
conventional concrete, they may give entirely different predictions for the same mixture of
concrete. In addition, Chapter Three presents existing models from the literature to estimate the
modulus of elasticity, the modulus of rupture and the splitting tensile strength in conventional
concrete, lightweight concrete and self-compacting concrete.

The proposed models are compared and verified by the most frequently used models in the
design codes and experimental data. The results show a good agreement between the
experimental data and the predictions of the proposed models. Table (3.31) summarizes the
developed models for the different mechanical properties of structural EPS-LWC. The models
for non-structural EPS-LWC are presented in this chapter also.

Table 3.30: Proposed models for structural EPS lightweight concrete

Parameter Equation

Modulus of elasticity (GPa, MPa) ‫ܧ‬௖ ൌ ͶǤʹͳሺ݂௖ᇱ ሻ଴Ǥହ଴଼

Modulus of rupture (MPa) ݂௥ ൌ ͳǤʹ͸͸Ǥ ሺ݂௖ᇱ ሻ଴Ǥସ଴଺ଶ

Splitting tensile strength (MPa) ݂௖௧ ൌ ͲǤͳͺ͸ʹሺ݂௖ᇱ ሻ଴Ǥ଼଺ହହ

According to the topic of this research program, a broad range of the existing models to
predict the mechanical properties of conventional concrete and self-compacting concrete are
also presented and evaluated in this chapter.

111
Chapter 3: Hardened concrete properties

REFERENCES

1. ASTM-C578, Standard specification for rigid, cellular polystyrene thermal insulation.


American society for Testing Materials (ASTM), C16.22, 2015. 04.06.
2. AS-1366.3, Rigid cellular plastics sheets for thermal insulation PART 3: Rigid cellular
polystyrene Moulded (RC/PS—M). Australian Standards, 1992.
3. Bagon, C. and S. Frondistou-Yannas, Marine floating concrete made with polystyrene expanded
beads. Magazine of Concrete Research, 1976. 28(97): p. 225-229.
4. Bischoff, P., K. Yamura, and S. Perry. Polystyrene aggregate concrete subjected to hard impact.
in ICE Proceedings. 1990. Thomas Telford.
5. Ravindrarajah, R.S. and A. Tuck, Properties of hardened concrete containing treated expanded
polystyrene beads. Cement and Concrete Composites, 1994. 16(4): p. 273-277.
6. Park, S.G. and D.H. Chisholm, Polystyrene Aggregate Concrete. STUDY REPORT, No. 85 ,
Building Research Levy, BRANZ, 1999.

7. Babu, K.G. and D.S. Babu, Behaviour of lightweight expanded polystyrene concrete containing
silica fume. Cement and Concrete Research, 2003. 33(5): p. 755-762.
8. Babu, K.G. and D.S. Babu, Performance of fly ash concretes containing lightweight EPS
aggregates. Cement and Concrete Composites, 2004. 26(6): p. 605-611.
9. Chen, B. and J. Liu, Properties of lightweight expanded polystyrene concrete reinforced with
steel fiber. Cement and Concrete Research, 2004. 34(7): p. 1259-1263.
10. Miled, K., et al., Compressive behavior of an idealized EPS lightweight concrete: size effects
and failure mode. Mechanics of materials, 2004. 36(11): p. 1031-1046.
11. Le Roy, R., E. Parant, and C. Boulay, Taking into account the inclusions' size in lightweight
concrete compressive strength prediction. Cement and Concrete Research, 2005. 35(4): p. 770-
775.
12. Babu, D.S., K.G. Babu, and T. Wee, Properties of lightweight expanded polystyrene aggregate
concretes containing fly ash. Cement and Concrete Research, 2005. 35(6): p. 1218-1223.
13. Chen, B. and J. Liu, Mechanical properties of polymer-modified concretes containing expanded
polystyrene beads. Construction and Building Materials, 2005. xxx: p. 1-5.

14. Babu, D.S., K.G. Babu, and W. Tiong-Huan, Effect of polystyrene aggregate size on strength
and moisture migration characteristics of lightweight concrete. Cement and Concrete
Composites, 2006. 28(6): p. 520-527.
15. Haghi, A., M. Arabani, and H. Ahmadi, Applications of expanded polystyrene (EPS) beads and
polyamide-66 in civil engineering, Part One: Lightweight polymeric concrete. Composite
Interfaces, 2006. 13(4-6): p. 441-450.
16. Bouvard, D., et al., Characterization and simulation of microstructure and properties of EPS
lightweight concrete. Cement and Concrete Research, 2007. 37(12): p. 1666-1673.
17. Chen, B. and J. Liu, Mechanical properties of polymer-modified concretes containing expanded
polystyrene beads. Construction and Building Materials, 2007. 21(1): p. 7-11.
18. Miled, K., K. Sab, and R. Le Roy, Particle size effect on EPS lightweight concrete compressive
strength: experimental investigation and modelling. Mechanics of Materials, 2007. 39(3): p.
222-240.
19. Amianti, M. and V.R. Botaro, Recycling of EPS: A new methodology for production of concrete
impregnated with polystyrene (CIP). Cement and Concrete Composites, 2008. 30(1): p. 23-28.
20. Bisschop, J. and J.G. van Mier, Effect of aggregates and microcracks on the drying rate of
cementitious composites. Cement and Concrete Research, 2008. 38(10): p. 1190-1196.
21. Lepech, M.D., et al., Design of green engineered cementitious composites for improved
sustainability. ACI Materials Journal, 2008. 105(6): p. 567-575.
22. Chen, B. and C. Fang, Contribution of fibres to the properties of EPS lightweight concrete.
Magazine of concrete research, 2009. 61(9): p. 671-678.
23. Bai, E.L., J.Y. Xu, and Z.G. Gao. Study on Deformation Property of EPS Concrete Under
Impact Loading. in Applied Mechanics and Materials. 2011. Trans Tech Publ.
24. Chen, B. and C. Fang, Mechanical properties of EPS lightweight concrete. Proceedings of the
ICE-Construction Materials, 2011. 164(4): p. 173-180.
25. Ling, I. and D. Teo, Properties of EPS RHA lightweight concrete bricks under different curing
conditions. Construction and Building Materials, 2011. 25(8): p. 3648-3655.

112
Chapter 3: Hardened concrete properties

26. Madandoust, R., M.M. Ranjbar, and S.Y. Mousavi, An investigation on the fresh properties of
self-compacted lightweight concrete containing expanded polystyrene. Construction and
Building Materials, 2011. 25(9): p. 3721-3731.
27. Miled, K., K. Sab, and R. Le Roy, Effective elastic properties of porous materials:
Homogenization schemes vs experimental data. Mechanics Research Communications, 2011.
38(2): p. 131-135.
28. Sadrmomtazi, A., et al., Properties of multi-strength grade EPS concrete containing silica fume
and rice husk ash. Construction and Building Materials, 2012. 35: p. 211-219.
29. Xu, Y., et al., Mechanical properties of expanded polystyrene lightweight aggregate concrete
and brick. Construction and Building Materials, 2012. 27(1): p. 32-38.
30. Carbonari, G., et al., Flexural behaviour of light-weight sandwich panels composed by concrete
and EPS. Construction and Building Materials, 2012. 35: p. 792-799.
31. Ferrándiz-Mas, V. and E. García-Alcocel, Physical and mechanical characterization of Portland
cement mortars made with expanded polystyrene particles addition (EPS). Materiales de
construccion, 2012. 62(308): p. 547-566.
32. Liu, A.F. and S.R. Guo. Research on Application of New Energy-Saving Building Materials. in
Applied Mechanics and Materials. 2012. Trans Tech Publ.
33. Passa, D.S., et al. Thermal and drying cyclic loading for cement based mortars and expanded
polystyrene foam layers. in Applied Mechanics and Materials. 2012. Trans Tech Publ.
34. Zhao, D.H., J.Y. Xu, and E.L. Bai. Study on Toughening Effect in EPS Concrete. in Applied
Mechanics and Materials. 2012. Trans Tech Publ.
35. Trussoni, M., C.D. Hays, and R.F. Zollo, Fracture Properties of Concrete Containing Expanded
Polystyrene Aggregate Replacement. ACI Materials Journal, 2013. 110(5).
36. Wang, H.Z. and D.G. Fu. Experimental study of porous concrete material thermal insulation
material lightweight polystyrene particles. in Applied Mechanics and Materials. 2013. Trans
Tech Publ.
37. Chen, B. and N. Liu, A novel lightweight concrete-fabrication and its thermal and mechanical
properties. Construction and Building Materials, 2013. 44: p. 691-698.
38. Elsalah, J., et al. The influence of recycled expanded polystyrene (EPS) on concrete properties:
Influence on flexural strength, water absorption and shrinkage. in 3RD INTERNATIONAL
ADVANCES IN APPLIED PHYSICS AND MATERIALS SCIENCE CONGRESS. 2013. AIP
Publishing.
39. Ferrándiz-Mas, V. and E. García-Alcocel, Durability of expanded polystyrene mortars.
Construction and Building Materials, 2013. 46: p. 175-182.
40. Herki, A., J. Khatib, and E. Negim, Lightweight concrete made from waste polystyrene and fly
ash. World Applied Sciences Journal, 2013. 21(9): p. 1356-1360.
41. Sadrmomtazi, A., J. Sobhani, and M. Mirgozar, Modeling compressive strength of EPS
lightweight concrete using regression, neural network and ANFIS. Construction and Building
Materials, 2013. 42: p. 205-216.
42. Hongbo, Z., et al., Impact resistance of a novel expanded polystyrene cement-based material.
Journal of Wuhan University of Technology-Mater. Sci. Ed., 2014. 29(2): p. 284-290.
43. Hidayat, I. and Gunaedi1, The Influence of Fly Ash in Concrete Compressive Strengthby using
Expanded Polystyrene as a Partial Substitution of Sand. International Journal of Applied
Engineering Research, 2014. 9(21): p. 9729-9738.

44. Kekanović, M., et al., Lightweight concrete with recycled ground expanded polystyrene
aggregate. Tehnički vjesnik, 2014. 21(2): p. 309-315.
45. Liu, N. and B. Chen, Experimental study of the influence of EPS particle size on the mechanical
properties of EPS lightweight concrete. Construction and Building Materials, 2014. 68: p. 227-
232.
46. Mousavi, S.A., S.M. Zahrai, and A. Bahrami-Rad, Quasi-static cyclic tests on super-lightweight
EPS concrete shear walls. Engineering Structures, 2014. 65: p. 62-75.
47. Schackow, A., et al., Mechanical and thermal properties of lightweight concretes with
vermiculite and EPS using air-entraining agent. Construction and Building Materials, 2014. 57:
p. 190-197.
48. Tamut, T., et al., PARTIAL REPLACEMENT OF COARSE AGGREGATES BY EXPANDED
POLYSTYRENE BEADS IN CONCRETE. International Journal of Research in Engineering and
Technology, 2014. 3(2): p. 238-241.
49. Nekooie, M.A., H. Goodall, and M.I. Mohamad, Selection of sustainable materials for
lightweight concrete. 2015.
113
Chapter 3: Hardened concrete properties

50. Xu, Y., et al., Prediction of compressive strength and elastic modulus of expanded polystyrene
lightweight concrete. 2015.
51. Li, Y., N. Liu, and B. Chen, Properties of lightweight concrete composed of magnesia phosphate
cement and expanded polystyrene aggregates. Materials and Structures, 2015. 48(1-2): p. 269-
276.
52. Lo Monte, F., P. Bamonte, and P.G. Gambarova, Physical and mechanical properties of heatǦ
damaged structural concrete containing expanded polystyrene syntherized particles. Fire and
Materials, 2015. 39(1): p. 58-71.
53. Pecce, M., et al., Steel–concrete bond behaviour of lightweight concrete with expanded
polystyrene (EPS). Materials and Structures, 2015. 48(1-2): p. 139-152.
54. Ranjbar, M.M. and S.Y. Mousavi, Strength and durability assessment of self-compacted
lightweight concrete containing expanded polystyrene. Materials and Structures, 2015. 48(4): p.
1001-1011.
55. Wang, C. and D. Arellano. Are the Mechanical Properties of Recycled-Content Expanded
Polystyrene (EPS) Comparable to Nonrecycled EPS Geofoam? in Geo-Congress 2014 Technical
Papers@ sGeo-characterization and Modeling for Sustainability. 2014. ASCE.
56. Gao, T., et al., Aerogel-incorporated concrete: An experimental study. Construction and
Building Materials, 2014. 52: p. 130-136.
57. AASHTO, LRFD, Bridge Design Specifications. Fourth Edition, American Association of State
Highway and Transportation Officials, Washington, D.C., 2007.
58. Rizkalla, S., Application of the LRFD bridge design specifications to high-strength structural
concrete: Flexure and compression provisions. Vol. 595. 2007: Transportation Research Board.
59. ACI-318-08, Building code requirements for structural concrete. American Concrete Institute,
International Organization for Standardization, Farmington Hills MI): American Concrete
Institute, 2008.
60. ACI-318-99, Building code requirements for structural concrete. ACI Committee, Standard ACI
318R–95. American Concrete Institute, Farmington Hills, Mich, 1999.
61. ACI-209-2R, Guide for Modeling and Calculating Shrinkage and Creep in Hardened Concrete.
Reported by ACI Committee 209, 2008.
62. 312-92, A.-. Building code requirements for reinforced concrete. ACI manual of concrete
practice, part 3, 1994.
63. CSA.A23.3-04, Design of concrete structures. Canadian Standards Association (CSA), 2010.
64. NZS-3101, Concrete Structures Standard, Part 1- The Design of Concrete Structures. New
Zealand Standard, Standards New Zealand, Wellington, New Zealand, 2006.
65. AS-3600-09, Concrete structures. Australian Standard, 2009.
66. NS.3473, Concrete structures-design rules. Norwegian Council for Strandardization, Norges
Standardiserings Forbund, 1992.
67. AIJ, Standard for Structural calculations of steel reinforced concrete structures. Architectural
Institute of Japan, 1985.
68. Yusuf, I. and Y. Jimoh, The transfer models of compressive to tensile, flexural and elastic
properties of palm kernel shell concrete. International Journal of Engineering, IJE, 2013. 11(2):
p. 195-200.
69. JSCE, Guidelines for concrete, Standard specifications for concrete structures- Design. Japan
Society of Civil Engineering, JSCE, 2010.
70. RakMK-D3-12, Rakennusten energiatehokkuus. Määräykset ja ohjeet Finnish building code ,
Suomen rakentamismääräyskokoelma, Helsinki, Ympäristöministeriö, 2012.
71. NTE-E.060, Reglamento nacional de edificaciones. Norma Técnica de Edificaciones E.060
Concreto Armado,Perú, 2009.
72. BS-8110-1, Structural use of concrete, Part14: Code of practice for design and construction.
Committee reference B/525/2 Draft for comment 95/105430 DC, 1997.
73. BS-5400-4, Steel, concrete and composite bridges, Part 4: Code of practice for design of
concrete bridges. Committee reference CSB/59, Draft for comment 88/11190 DC , British
Standard, 1999.
74. SP-52-101, Concrete and reinforced concrete structures without prestressing. Russian standard,
Code of Practice in building construction (SP), 2003.
75. Carrasquillo, R.L., H.H. Nilson, and F.O. Slate, Properties of high strength concrete subjected to
short term loads. ACI J.Proc, 1981. 78(3): p. 171-178.
76. Soleymani, H.R., Structural design properties of concrete for a bridge in Alberta. Canadian
Journal of Civil Engineering, 2006. 33(2): p. 199-205.

114
Chapter 3: Hardened concrete properties

77. Haranki, B., Strength, modulus of elasticity, creep and shrinkage of concrete used in florida.
2009, University of Florida.
78. Ahmad, S.H. and S.P. Shah, Structural properties of high-strength concrete and its implications
for precast concrete. 1985.
79. Jobse, H. and S. Moustafa, Applications of high strength concrete for highway bridges.
PRECAST/PRESTRESSED CONCRETE INSTITUTE. JOURNAL, 1984. 29(3).
80. Cook, J.E., 10,000 psi Concrete. Concrete International, 1989. 11(10): p. 67-75.
81. Zhang, M.H. and O.E. Gjvorv, Mechanical properties of high-strength lightweight concrete.
ACI Materials Journal, 1991. 88(3).
82. Jensen, V. The Plasticity Ratio of Concrete and It's Effect on the Ultimate Strength of Beamse. in
ACI Journal Proceedings. 1943. ACI.
83. Pauw, A. Static modulus of elasticity of concrete as affected by density. in ACI Journal
Proceedings. 1960. ACI.
84. Floyd 0. Slate, A.H.N. and S. Martinez. Mechanical properties of high-strength lightweight
concrete. in ACI Journal Proceedings. 1986. ACI.
85. Morales, S.M., Short-term mechanical properties of high-strength light-weight concrete. NASA
STI/Recon Technical Report N, 1982. 84: p. 10387.
86. Meyer, K.F. and L.F. Kahn, Lightweight concrete reduces weight and increases span length of
pretensioned concrete bridge girders. PCI journal, 2002. 47(1): p. 68-75.
87. Alengaram, U.J., H. Mahmud, and M.Z. Jumaat, Enhancement and prediction of modulus of
elasticity of palm kernel shell concrete. Materials & Design, 2011. 32(4): p. 2143-2148.
88. Ahmmad, R., et al., Ductility performance of lightweight concrete element containing massive
palm shell clinker. Construction and Building Materials, 2014. 63: p. 234-241.
89. Dinakar, P., Properties of fly-ash lightweight aggregate concretes. Proceedings of the ICE-
Construction Materials, 2012. 166(3): p. 133-140.
90. Tassew, S.T. and A.S. Lubell, Mechanical properties of lightweight ceramic concrete. Materials
and structures, 2012. 45(4): p. 561-574.
91. Cousins, T.E., C.L. Roberts-Wollmann, and M.C. Brown, High-performance/high-strength
lightweight concrete for bridge girders and decks. Vol. 733. 2013: Transportation Research
Board.
92. Alengaram, U.J., B.A. Al Muhit, and M.Z. bin Jumaat, Utilization of oil palm kernel shell as
lightweight aggregate in concrete–a review. Construction and Building Materials, 2013. 38: p.
161-172.
93. 213R-03, A., Guide for Structural Lightweight-Aggregate Concrete. American Concrete
Institute, ACI Committee 213, 2003.
94. FHWA, Lightweight Concrete: Mechanical Properties. Federal Highway Administration
(FHWA), Highway Research Center, FHWA Publication No. FHWA-HRT-13-061, 2013.
95. AASHTO, AASHTO LRFD Bridge Design Specifications. American Association of State
Highway and Transportation Officials 2012. 6th Edition.
96. Bogas, J.A. and R. Nogueira, Tensile strength of structural expanded clay lightweight concrete
subjected to different curing conditions. KSCE Journal of Civil Engineering, 2014. 18(6): p.
1780-1791.
97. Shafigh, P., et al., Engineering properties of oil palm shell lightweight concrete containing fly
ash. Materials & Design, 2013. 49: p. 613-621.
98. (EC-2), E., Design of Concrete Structures. Part 1-General Rules and Rules for Buildings.
German version EN 1992-1-1:2004 + AC: 2010, The Concrete Centre: Blackwater, Camberley,
UK, , 2002.
99. Wegian, F., Strength properties of lightweight concrete made with LECA grading. Australian
Journal of Civil Engineering, 2012. 10(1): p. 11.
100. Yew, M.K., et al., Effects of Low Volume Fraction of Polyvinyl Alcohol Fibers on the
Mechanical Properties of Oil Palm Shell Lightweight Concrete. Advances in Materials Science
and Engineering, 2014.
101. Kou, S., et al., Properties of lightweight aggregate concrete prepared with PVC granules
derived from scraped PVC pipes. Waste Management, 2009. 29(2): p. 621-628.
102. OHBDC, Ontario Highway Bridge Design Code. Ministry of Transportation, Downs view,
Ontario, Canada, 2nd Edition, 1983.
103. CEB-FIP, High-strength concrete state of the art report. London: Thomas Telford, 1990.
104. Beton, C.E.-I.d., CEB-FIP model code 1990. London. Thomas Telford, 1993.
105. GBJ-11-89, National standard of the People’s Republic of China, Chinese code for seismic
design of buildings. Chinese Design Code, Beijing: NewWorld Press, 1994.
115
Chapter 3: Hardened concrete properties

106. ACI363-R, State of the art report on high-strength concrete. Farmington Hills (Michigan):
American Concrete Institute,, 1992.
107. EHE, Spanish code for structural concrete. Real Decreto 2661/1998, Madrid; [in Spanish], 1998.
108. FHWA, Material Property Characterization of Ultra-High Performance Concrete. Federal
Highway Administration (FHWA), Highway Research Center, McLean, VA 22101-2296, 2000.
109. TS-500, Building code requirements for reinforced concrete. Turkish Standards Institute,
Ankara, Turkey, 2000.
110. IS.456, Plain and Reinforced concrete- code of practice. Indian Standard, ICS, 91.100.30, New
Delhi, July 2000.
111. GDC-2000, Greek Design Code. Greek National Code Center, Athens, Greece, 2000.
112. NCHRP, National Cooperative Highway Research Program. 2003.
113. NBR.6118, Design of concrete structures. Brazilian association of technical standards , Rio de
Janeiro; [in Portuguese], 2003.
114. IDC-3274, Come Modificato Dall’OPCM 3431. Italian Design Code, Italy: Public and Health
Ministry, 2003.
115. Fanourakis, G. and Y. Ballim, An assessment of the accuracy of nine design models for
predicting creep in concrete: technical paper. Journal of the South African Institution of Civil
Engineering= Joernaal van die Suid-Afrikaanse Instituut van Siviele Ingenieurswese, 2006.
48(4): p. 2-8.
116. Ravindrarajah, R.S. and C. Tam, Properties of concrete made with crushed concrete as coarse
aggregate. Magazine of Concrete Research, 1985. 37(130): p. 29-38.
117. Gardner, N. and J. Zhao. Mechanical properties of concrete for calculation of long term
deformations. in Proceedings of the Second Canadian Symposium on Cement and Concrete.
1991. University of British Columbia Press, Vancouver, British Columbia, Canada.
118. Gutierrez, P.A. and M.F. Canovas, The modulus of elasticity of high performance concrete.
Materials and Structures, 1995. 28(10): p. 559-568.
119. Rashid, M., M. Mansur, and P. Paramasivam, Correlations between mechanical properties of
high-strength concrete. Journal of Materials in Civil Engineering, 2002. 14(3): p. 230-238.
120. Levtchitch, V., et al. Seismic Performance Capacities of Old Concrete. in 13 WCEE: 13 th
World Conference on Earthquake Engineering Conference Proceedings. 2004.
121. Leemann, A. and C. Hoffmann, Properties of self-compacting and conventional concrete–
differences and similarities. Magazine of Concrete Research, 2005. 57(6): p. 315-319.
122. Kheder, G. and S. Al-Windawi, Variation in mechanical properties of natural and recycled
aggregate concrete as related to the strength of their binding mortar. Materials and Structures,
2005. 38(7): p. 701-709.
123. Liu, Y., Strength, modulus of elasticity, shrinkage and creep of concrete. 2007: ProQuest.
124. Dinakar, P., K. Babu, and M. Santhanam, Mechanical properties of high-volume fly ash self-
compacting concrete mixtures. Structural concrete, 2008. 9(2): p. 109-116.
125. Noguchi, T., et al., A practical equation for elastic modulus of concrete. ACI Structural Journal,
2009. 106(5).
126. Obispo, S.L., Shrinkage and modulus of elasticity in concrete with recycled aggregates. Master
thesis, California Polytechnic State University,, 2011.
127. Aslani, F. and S. Nejadi, Mechanical properties of conventional and self-compacting concrete:
An analytical study. Construction and Building Materials, 2012. 36(0): p. 330-347.
128. EC-2, E., Design of Concrete Structures, Part 1: General Rules and Rules for buildings.
Commission of European Communities ENV, German version EN 1992-1-1:2004 + AC: 2010,
The Concrete Centre: Blackwater, Camberley, UK, . 2004: p. 1-1.
129. ACI-363-R, State of the art report on high-strength concrete. armington Hills (Michigan):
American Concrete Institute (ACI), 1992.
130. IS.456, Plain and Reinforced concrete- code of practice. Indian Standard, ICS, 91.100.30, New
Delhi, 2000.
131. ACI-435, Deflections of reinforced concrete flexural member. American Concrete Institute
(ACI), Concrete International, 1968.
132. Carrasquilio, R.L. and A.H. Nilson. Properties of high strength concrete subject to short-term
loads. in ACI Journal Proceedings. 1981. ACI.
133. Bakhsh, A.H., F.F. Wafa, and A.A. Akhtaruzzaman, Torsional behavior of plain high-strength
concrete beams. ACI Structural Journal, 1990. 87(5).
134. Raphael, J.M. Tensile strength of concrete. in ACI Journal Proceedings. 1984. ACI.
135. Lane, D.S., Evaluation of concrete characteristics for rigid pavements. 1998.
136. Akazawa, T., Tension test method for concrete. Bulletin, 1953. 16.
116
Chapter 3: Hardened concrete properties

137. Carneiro, F. and A. Barcellos, Concrete tensile strength. RILEM Bulletin, 1953. 13: p. 103-107.
138. Vinayaka, M., The relation between crushing strength and tensile splitting strength of concrete.
CBJ. 29th Session, Hyderabad, 1959.
139. Sen, B. and P. Desayi, Determination of tensile strength of concrete by splitting a cube along its
diagonal plane. Indian Concrete Journal, 1962. 36(7): p. 249-252.
140. Carino, N.J. and H. Lew. Re-examination of the relation between splitting tensile and
compressive strength of normal weight concrete. in ACI Journal Proceedings. 1982. ACI.
141. Gardner, N., Relationship of the punching shear capacity of reinforced concrete slabs with
concrete strength. ACI Structural Journal, 1990. 87(1).
142. Oluokun, F., Prediction of concrete tensile strength from its compressive strength: an evaluation
of existing relations for normal weight concrete. ACI Materials Journal, 1991. 88(3).
143. Burg, R.G. and B.W. Ost, Engineering properties of commercially available high-strength
concretes. 1992.
144. Beton, C.E.-I.d., CEB-FIP model code 1990. London, Thomas Telford, 1993.
145. NEN-6722, Regulations concrete. Construction (VBU 1988), with correction sheet; 2000, 2000.
146. Hueste, M.B.D., et al., Mechanical properties of high-strength concrete for prestressed
members. ACI Structural Journal, 2004. 101(4).
147. AASHTO-07, Specifications, LRFD Bridge Design. American Association of State Highway and
Transportation Officials, Washington DC, 2007. 20001.
148. NS.3473, Concrete structures-design rules. Norwegian Council for Strandardization, Norges
Standardiserings Forbund., 1992.
149. Persson, B., A comparison between mechanical properties of self-compacting concrete and the
corresponding properties of normal concrete. Cement and concrete Research, 2001. 31(2): p.
193-198.
150. Felekoğlu, B., S. Türkel, and B. Baradan, Effect of water/cement ratio on the fresh and hardened
properties of self-compacting concrete. Building and Environment, 2007. 42(4): p. 1795-1802.
151. Atahan, H., D. Trejo, and M. Hueste, Applicability of standard equations for predicting
mechanical properties of SCC. ACI Special Publication, 2007. 247.
152. Kim, Y.H., Characterization of self-consolidating concrete for the design of precast,
pretensioned bridge superstructure elements. 2008: TEXAS A&M UNIVERSITY.
153. ACI-318-08. Building code requirements for structural concrete (ACI 318-08) and commentary.
2008. American Concrete Institute, International Organization for Standardization.
154. Vilanova, A., J. Fernandez-Gomez, and G.A. Landsberger, Evaluation of the mechanical
properties of self compacting concrete using current estimating models: Estimating the modulus
of elasticity, tensile strength, and modulus of rupture of self compacting concrete. Construction
and Building Materials, 2011. 25(8): p. 3417-3426.
155. Yıldırım, H. and O. Sengul, Modulus of elasticity of substandard and normal concretes.
Construction and Building Materials, 2011. 25(4): p. 1645-1652.
156. Das, D. and A. Chatterjee, A comparison of hardened properties of fly-ash-based self-
compacting concrete and normally compacted concrete under different curing conditions.
Magazine of Concrete Research, 2011. 64(2): p. 129-141.
157. Hu, W., et al., Constitutive models for regression of various experimental stress–strain relations.
International Journal of Mechanical Sciences, 2015. 101: p. 1-9.
158. Sukumar, B., K. Nagamani, and R.S. Raghavan, Evaluation of strength at early ages of self-
compacting concrete with high volume fly ash. Construction and Building Materials, 2008.
22(7): p. 1394-1401.
159. Topçu, İ.B. and T. Uygunoğlu, Effect of aggregate type on properties of hardened self-
consolidating lightweight concrete (SCLC). Construction and Building Materials, 2010. 24(7): p.
1286-1295.
160. Parra, C., M. Valcuende, and F. Gómez, Splitting tensile strength and modulus of elasticity of
self-compacting concrete. Construction and Building materials, 2011. 25(1): p. 201-207.
161. Turatsinze, A. and M. Garros, On the modulus of elasticity and strain capacity of self-
compacting concrete incorporating rubber aggregates. Resources, conservation and recycling,
2008. 52(10): p. 1209-1215.
162. Tasnimi, A.A. Mathematical model for complete stress&#8211;strain curve prediction of
normal, light-weight and high-strength concretes. Magazine of Concrete Research, 2004. 56, 23-
34.
163. Almusallam, T. and S. Alsayed, Stress-strain relationship of normal, hight and lightweight
concrete. Magazine of Concrete Research, 1995. 47(170): p. 39-44.

117
Chapter 3: Hardened concrete properties

164. Carreira, D.C., KH Stress-Strain Relationship for plain concrete in compression Journal of the
American Concrete Institute, 1985. 82(6): p. 797-804.
165. Tasnimi, A.A., Mathematical model for complete stress–strain curve prediction of normal, light-
weight and high-strength concretes. Magazine of Concrete Research, 2004. 56(1): p. 23-34.
166. Clarke, J.L., Structural light weight aggregate concrete. Blackie Academic and professional, an
imprint of Chapman and Hall, Wester Cleddens Road, Bishpbriggs, Glascow G642NZ, 1993.
167. Santiago, E.Q.R., et al., Mechanical behavior of recycled lightweight concrete using EVA waste
and CDW under moderate temperature. Ibracon structures and materials journal, 2009. 2(3): p.
211-221.
168. Richard, R.M. and B.J. Abbott, Versatile elastic-plastic stress-strain formula. Journal of the
Engineering Mechanics Division, 1975. 101(4): p. 511-515.
169. Wang, P., S. Shah, and A. Naaman. Stress-strain curves of normal and lightweight concrete in
compression. in ACI Journal Proceedings. 1978. ACI.
170. Hussin, M., et al. A mathematical model for complete stress-strain curve prediction of permeable
concrete. in Proceedings of the 22nd Australasian Conference on the Mechanics of Structures
and Materials (ACMSM 22). 2013. Taylor & Francis (CRC Press)/Balkema.
171. Hognestad, E., Study of combined bending and axial load in reinforced concrete members.
University of Illinois. Engineering Experiment Station. Bulletin; no. 399, 1951.
172. Smith, G. and L. Young. Ultimate theory in flexure by exponential function. in ACI Journal
Proceedings. 1955. ACI.
173. Desayi, P. and S. Krishnan. Equation for the stress-strain curve of concrete. in ACI Journal
Proceedings. 1964. ACI.
174. Saenz, L., Equation for the stress–strain curve of concrete (Discussion). J ACI, 1964. 61(9): p.
1229–35.
175. Mazars, J. and G. Pijaudier-Cabot, Continuum damage theory-application to concrete. Journal of
Engineering Mechanics, 1989. 115(2): p. 345-365.
176. Collins, M.P. and D. Mitchell, Prestressed concrete structures. Vol. 9. 1991: Prentice Hall
Englewood Cliffs, NJ.
177. Wee, T., M. Chin, and M. Mansur, Stress-strain relationship of high-strength concrete in
compression. Journal of Materials in Civil Engineering, 1996. 8(2): p. 70-76.
178. Chandrasekhar, M., M.S. Rao, and M. Janardhana, Studies on stress-strain behavior of SFRSCC
and GFRSCC under axial compression. International Journal of Earth Sciences and Engineering,
ISSN, 2011: p. 0974-5904.
179. Prasad, M., P.R. Kumar, and T. Oshima, Development of Analytical Stress-Strain Model for
Glass Fiber Reinforced Self Compacting Concrete. International Journal of Mechanics & Solids,
2009. 4(1).
180. Kumar, R., B. Singh, and P. Bhargava, Flexural capacity predictions of self-compacting
concrete beams using stress–strain relationship in axial compression. Magazine of Concrete
Research, 2011. 63(1): p. 49-59.
181. Kim, D.-J., et al., Bond strength of steel deformed rebars embedded in artificial lightweight
aggregate concrete. Journal of Adhesion Science and Technology, 2013. 27(5-6): p. 490-507.
182. Orangun, C., J. Jirsa, and J. Breen. A reevaulation of test data on development length and
splices. in ACI Journal Proceedings. 1977. ACI.
183. Kemp, E. and W. Wilhelm. Investigation of the parameters influencing bond cracking. in ACI
Journal Proceedings. 1979. ACI.
184. Eligehausen, R., Übergreifungsstöße zugbeanspruchter Rippenstäbe mit geraden Stabenden.
Deutscher Ausschuss für Stahlbeton, 1979(301).
185. Kemp, E.L. Bond in reinforced concrete: behavior and design criteria. in ACI Journal
Proceedings. 1986. ACI.
186. Robins, P. and S. Austin, Bond of lightweight aggregate concrete incorporating condensed silica
fume. ACI Special Publication, 1986. 91.
187. Chapman, R.A. and S.P. Shah, Early-age bond strength in reinforced concrete. ACI Materials
Journal, 1987. 84(6).
188. Harajli, M., Development/splice strength of reinforcing bars embedded in plain and fiber
reinforced concrete. ACI structural Journal, 1994. 91(5).
189. Huang, Z., B. Engström, and J. Magnusson, Experimental investigation of the bond and
anchorage behaviour of deformed bars in high strength concrete. Goteborg: Division of
Concrete Structures, Chalmers University of Technology, 1996.
190. Esfahani, M.R. and B.V. Rangan, Bond between normal strength and high-strength concrete
(HSC) and reinforcing bars in splices in beams. ACI Structural Journal, 1998. 95(3).
118
Chapter 3: Hardened concrete properties

191. Pillai, S.U., D.W. Kirk, and M.-A. Erki, Reinforced concrete design. 1988: McGraw-Hill
Ryerson.
192. Zuo, J. and D. Darwin, Splice strength of conventional and high relative rib area bars in normal
and high-strength concrete. ACI structural Journal, 2000. 97(4).
193. Committee, A., Bond and Development of Straight Reinforcing Bars in Tension (ACI 408R-03).
American Concrete Institute, Detroit, Michigan, US, 2003.
194. Sancak, E., Prediction of bond strength of lightweight concretes by using artificial neural
networks. Sci. Res. Essays, 2009. 4(4): p. 256-266.
195. Tanyildizi, H., Fuzzy logic model for the prediction of bond strength of high-strength lightweight
concrete. Advances in Engineering Software, 2009. 40(3): p. 161-169.
196. Desnerck, P., G. De Schutter, and L. Taerwe, Bond behaviour of reinforcing bars in self-
compacting concrete: experimental determination by using beam tests. Materials and Structures,
2010. 43(1): p. 53-62.
197. Aslani, F. and S. Nejadi, Bond behavior of reinforcement in conventional and self-compacting
concrete. Advances in Structural Engineering, 2012. 15(12): p. 2033-2052.
198. Nilson, S., Bond Stress-Slip Relationships in Reinforced Concrete. DepL of Structure
Engineering, Comell Univ., Ithaca, NY, 1971.
199. Wu, Y.-F. and X.-M. Zhao, Unified bond stress–slip model for reinforced concrete. Journal of
Structural Engineering, 2012. 139(11): p. 1951-1962.
200. Diab, A.M., et al., Bond behavior and assessment of design ultimate bond stress of normal and
high strength concrete. Alexandria Engineering Journal, 2014. 53(2): p. 355-371.
201. Marti, P., et al., Tension chord model for structural concrete. Structural Engineering
International, 1998. 8(4): p. 287-298.
202. Oh, B.H. and S.H. Kim, Realistic models for local bond stress-slip of reinforced concrete under
repeated loading. Journal of Structural Engineering, 2007. 133(2): p. 216-224.
203. Bae, S. Mix design, formwork pressure and bond characteristics of special self-consolidating
concrete. in Masters Abstracts International. 2006.

119
Chapter 4 – Time-dependent behaviour of concrete

CHAPTER 4

TIME-DEPENDENT BEHAVIOUR OF
CONCRETE
Chapter 4 – Time-dependent behaviour of concrete

Table of Contents

4.1. INTRODUCTION ...................................................................................................................... 120


4.2. CREEP STRAIN ......................................................................................................................... 123
4.3. CREEP COEFFICIENT .............................................................................................................. 125
4.4. SHRINKAGE STRAIN .............................................................................................................. 126
4.5. EXISTING CREEP AND SHRINKAGE MODELS FOR LIGHTWEIGHT, SELF-
COMPACTING AND CONVENTIONAL CONCRETE ....................................................................... 127
4.5.1. Creep and Shrinkage Models for Lightweight Concrete .................................................... 127
4.5.1.1. Creep models for lightweight concrete ...................................................................... 127
4.5.1.2. Shrinkage models of lightweight concrete ................................................................. 130
4.5.2. Creep and Shrinkage Models of Conventional Concrete.................................................... 130
4.5.2.1. Creep models for conventional concrete ................................................................... 130
4.5.2.2. Shrinkage models for conventional concrete ............................................................. 133
4.5.3. Creep and Shrinkage Models in Self-compacting Concrete ............................................... 134
4.5.3.1. Creep models of self-compacting concrete ................................................................ 134
4.5.3.2. Shrinkage models of self-compacting concrete ......................................................... 135
4.6. SOME EXPERIMENTS ON CREEP AND SHRINKAGE OF LIGHTWEIGHT CONCRETE IN
THE LITERATURE ................................................................................................................................ 136
4.7. BOND TRANSFER LENGTH IN LIGHTWEIGHT CONCRETE ........................................... 136
4.7.1. Time-dependent Bond Transfer Length ............................................................................. 138
4.7.2. Time-dependent Effects of Creep and Shrinkage on Bond Mechanism ............................. 140
4.7.3. Bond Behaviour After Cracking and Before Yielding of Steel .......................................... 141
4.7.4. Experimental Investigation on the Ratio of Long-term Bond Transfer Length to Initial Bond
Transfer Length in Conventional Concrete .......................................................................................... 143
4.7.5. Analytical Model of Bond Transfer Length ....................................................................... 144
4.8. EXISTING MODELS OF BOND-SLIP FOR LIGHTWEIGHT, CONVENTIONAL AND SELF-
COMPACTING CONCRETE ................................................................................................................. 149
4.8.1. Bond-Slip Models for Lightweight Concrete ..................................................................... 149
4.8.2. Bond-Slip Models in Conventional Concrete ..................................................................... 150
4.8.3. Bond-Slip Models in Self-compacting Concrete ................................................................ 153
4.9. SUMMARY ................................................................................................................................ 154
REFERENCES ......................................................................................................................................... 156
Chapter 4 – Time-dependent behaviour of concrete

List of Figures

Figure 4.1: General trend of shrinkage and creep strain in concrete under sustained loading .. 122
Figure 4.2: Compliance curves for various ages (t') at loading [11] ......................................... 123
Figure 4.3: Creep effects on the strain distribution of a singly reinforced cross-section [12] .. 125
Figure 4.4: Creep isochrones in concrete [11] .......................................................................... 126
Figure 4.5: Shrinkage-induced deformation and stresses in a singly reinforced beam [12] ..... 127
Figure 4.6: Direct and indirect tensile strength and modulus of rupture vs. compressive strength
[4] .............................................................................................................................................. 137
Figure 4.7: Bond transfer of tension reinforcement in concrete ............................................... 138
Figure 4.8: a) Tension stiffening model derived from uniaxial tension b) the equivalent concrete
stress-strain relationship [40] .................................................................................................... 139
Figure 4.9: Schematic deformation of concrete in a tension specimen [41] ............................. 140
Figure 4.10: Mechanism of cracking formation in a uniaxially reinforced concrete prism under
tension [30] ............................................................................................................................... 142
Figure 4.11: Schematic distribution of forces, strains, normal stresses, and bond stresses in
cracked member [47] ................................................................................................................ 142
Figure 4.12: Details of specimen in long-term bond transfer length test [48] .......................... 143
Figure 4.13: First cracking in a restrained direct tension member [51] .................................... 145
Figure 4.14: Variation of average stress in steel and concrete .................................................. 145
Figure 4.15: Bond-slip mechanisms in a) smooth bars; b) deformed bars [53] ........................ 146
Figure 5.16: Tension stiffening effect in strain-stress diagram respect to bond quality and steel
ratio [29] ................................................................................................................................... 147
Figure 4.17: Final concrete and steel stresses after direct tension cracking [51] ...................... 148
Figure 4.18: Normalized bond stress-slip for 25 and 35 mm bars under compression [56] ..... 150
Chapter 4 – Time-dependent behaviour of concrete

List of Tables

Table 4.1: Existing creep models of lightweight concrete ........................................................ 128


Table 4.2: Existing models to predict shrinkage strain in lightweight concrete ....................... 130
Table 4.3: Existing creep models of conventional concrete ..................................................... 131
Table 4.4: Existing shrinkage models for conventional concrete ............................................. 133
Table 4.5: Existing creep models for self-compacting concrete ............................................... 134
Table 4.6: Existing shrinkage models for self-compacting concrete ........................................ 135
Table 4.7: Details of test specimens (reinforcement and dimensions) ...................................... 144
Table 4.8: Material properties of concrete batch 1 (B.1) and batch 2 (B.2) ............................. 144
Table 4.9: Measured no-bond length after first and all shrinkage cracking .............................. 148
Table 4.10: Existing bond-slip models of light-weight concrete in literature ........................... 149
Table 4.11: Existing bond-slip models of conventional concrete in literature.......................... 151
Chapter 4 – Time-dependent behaviour of concrete

Chapter 4: Time-dependent properties of concrete

4.1. INTRODUCTION

Generally, when a concrete structure is subjected to an external loading, its response is


studied under instantaneous and time-dependent conditions separately. Regarding the time-
dependent behaviour of concrete structure, the following items could be addressed: [1]
- Time-dependent properties of concrete, especially creep and shrinkage, with their
dependence on a multiplicity of variables;
- Time-dependency of the concrete properties in regard to the processes of structural
analysis;
- The time-dependent stress and strain analysis of reinforced concrete sections;
- Integration of the sectional effects to determine the forces, deformations and deflections
of structural members.

There are a considerable number of investigations about the time-dependent behaviour of


concrete in the literature since it was first recognized and reported almost a century ago.
However, there is no distinctive boundary in Reinforced Concrete (RC) members as to when the
instantaneous (short-term) behaviour ends and long-term begins. In addition, there is no end for
time-dependent deflection of reinforced concrete structures during the service life that makes
the prediction of total deflection inaccurate.

To determine the long-term deflection of a reinforced concrete member, the time-delayed


effects of shrinkage, creep and the consequence reduction in tension stiffening need to be
accurately considered in the design. Additional factors which can contribute to the increased

120
Chapter 4 – Time-dependent behaviour of concrete

long-term deflection including the development of new cracks, widening of the earlier cracks,
and effects of the repeated load cycles should be addressed too [2].

Shrinkage induces strain that develops with time and causes additional curvature known as
shrinkage warping. Deflections also increase due to the effects of creep in the concrete member.
The tension stiffening effect is responsible for a significant proportion of concrete stiffness. It
decreases over time due to the effects of creep and shrinkage strains [3].

The time-dependent deflection also increases with increasing cracking of the cross-sections.
Cracking is caused when the tensile stress induced by flexure, shrinkage and thermal effects,
acting individually or simultaneously, becomes greater than the tensile strength of the concrete.
As the concrete drying shrinkage increases with time, cracking can also be expected to increase
with time. When a high level of concrete tensile stress is maintained for some time, more and
more cracks will occur. Therefore, an RC member that is essentially uncracked at the time of
stripping will gradually crack. It causes the effective moment of inertia to decrease, and this will
contribute to the long-term deflection [4]. Accordingly, a general equation to predict the long-
term deflection of concrete members can be written as Equation (4.1).

ο௧௢௧ ൌ ο௜௡௦ ൅  ο௦௛ ൅ ο௖௥ ൅ ο௧௠௣ Eq. (4. 1)

Where Δtot is total deflection, Δins is instantaneous deflection, Δsh is deflection due to
shrinkage (autogeneous and chemical shrinkage), Δcr is deflection due to creep effect and Δtmp is
temperature induced deflection.

The time-dependent effects of creep, shrinkage and temperature variation is the source of the
nonlinear behaviour of concrete structures that generally is not taken into account in
instantaneous deflection calculations of RC structures [5]

The creep and shrinkage characteristics of concrete are critical in design of any concrete
structural member which will be under a sustained load. These characteristics are essential,
since deformation of the member with time can be three or more times as great as the elastic
deformation occurring upon application of the load [6].

Despite the dependent nature of creep and shrinkage on each other [7], in most design
purposes they are assumed to be independent. Therefore, their individual effects are being added
to the calculation of the total strain in RC structure. Considering the effective parameters in
Equation (4.1), the total strain in the concrete may be considered as the summation of the strains
induced by all the affecting parameters.

121
Chapter 4 – Time-dependent behaviour of concrete

Generally, shrinkage is assumed to be a time-dependent parameter; however, creep is


changing by the applied stress level and time. Most of the investigations show that creep is very
sensitive to the curing process, environmental variations, and especially the concrete
composition [8].

Generally, varying deformation can be observed in an RC member subjected to a constant


sustained loading and constant temperature over time. This behaviour may be illustrated by the
effects of creep and shrinkage and the constant stress applied at time t0. The shrinkage consists
of two parts of autogeneous and drying shrinkage in general. After the concrete sets or moist
curing has finished, concrete undergoes autogeneous shrinkage (before drying) and subsequent
drying shrinkage. The shrinkage strain increases continuously at a decreasing rate.

However, creep includes basic creep and drying creep parts in general. Immediately after
applying the loading, drying creep and basic creep take place [9].

Figure (4.1) shows the terms relating to creep and shrinkage of concrete. It displays the
general trend and comparative magnitude of the creep and shrinkage components under constant
sustained load (stress).

In Figure (4.1), in contrast with the drying effects, the swelling portion of the diagram deals
with the water absorption and expansion of the concrete.

Figure 4.1: General trend of shrinkage and creep strain in concrete under sustained loading

122
Chapter 4 – Time-dependent behaviour of concrete

In real conditions, there are more complicated parameters affecting the time-dependent
deflection of concrete. For instance, changing of the stress history in the concrete member,
applied restraints and unstable environmental conditions e.g. temperature and humidity may
affect the deflection of a concrete structure. These uncertainties may strongly influence the
predicted values of creep and shrinkage induced strains for the time-dependent deflection
calculations.

4.2. CREEP STRAIN

Creep of concrete is the increase in strain under sustained loads. It is composed of two
components, the basic creep or deformation under the loads without moisture loss, and the
drying creep or deformation under drying conditions only. Creep of concrete is mainly due to
the creep of cement paste. For Portland cement, creep is primarily determined by the creep of
Calcium Silicate Hydrates (CSH) [10]. The time-dependent deformation due to creep is
attributed to the movement of water between different phases of concrete. When an external
load is applied, it changes the attraction forces between the cement gel particles. This
modification of forces causes an imbalance in the attractive and disjoining forces. However, the
imbalance is gradually eliminated by moisture transfer into the pores in cases of compression,
and away from the pores in instances of tension.

Creep influences both the strain (deflection) and stress distribution in the reinforced concrete
structures. Figure (4.2) shows the total load induced strain per unit stress caused by the unit of
uniaxial sustained load (also called as compliance function, J(t, t0)) applied at age t since
loading age t0.

t' = 5 days 50 days t' = 5 days 50 days


500 days 5000 days 500 days 5000 days
1/ E 1/ E
J(t, t')
J(t, t')

Time (t) 1 hour Time (t) 30 Years

Figure 4.2: Compliance curves for various ages (t') at loading [11]

123
Chapter 4 – Time-dependent behaviour of concrete

The gradual increase of creep strain in the compression zone of an RC member cross-section
increases the curvature and a consequent increase in deflection of the member. For a plain
concrete member without reinforcing bars, the strain development at every point on the section
is proportional to the creep coefficient, φ(t). Hence, it is proportional to the increase in curvature
as well.

Considering the cracked or uncracked RC section, the interaction between the reinforced
concrete section curvature and development of creep strain is graphically illustrated in Figure
(4.3). The creep strain influences the total strain distribution on the section and consequently the
curvature of the RC member.

Figure (4.3.a) shows the uncracked singly reinforced concrete section. The reinforcement in
tensile zone applies some restraint to the creep. Depending on the number and area of the tensile
steel bars, the increase in curvature due to creep is proportional to a large fraction of the creep
coefficient. Usually, this creep induced curvature is between 0.7 φ(t) and 0.95 φ(t) [12].

Figure (4.3.b) shows the fully cracked singly reinforced concrete section. Comparing with
Figure (4.3a), the initial curvature is considerably larger. Assuming that the cracked tensile
concrete below the neutral axis carries no stress; no creep will happen in that tensile concrete
zone. On the compression side, the creep reduces the neutral axis depth due to extension of the
crack, and consequently causes a dropping of the compressive stress level. The lower
compressive stress level reduces the creep strain, so the increase in curvature is proportional to a
small fraction of the creep coefficient. In a fully cracked RC section, the creep induced
curvature is usually less than 0.25 φ(t). Although the total deflection of the cracked flexural
member is significantly high, the creep effect in total deflection in an uncracked RC section is
relatively higher e.g. 0.25 φ(t) <0.7 φ(t) [12].

124
Chapter 4 – Time-dependent behaviour of concrete

(a)

(b)

Figure 4.3: Creep effects on the strain distribution of a singly reinforced cross-section [12]

4.3. CREEP COEFFICIENT

The ratio of creep strain to the initial strain or identically, the ratio of creep compliance to the
compliance obtained at early ages is called the creep coefficient [13]. It typically shows the
capacity of concrete to creep and is described by the general form of Equation (4.2).

Figure (4.4) exhibits the curves of stress versus creep strains for various constant load
durations. In the range of stress up to 40% - or 50% of the compressive strength (the typical
service stress range), the curves in Figure (4.4) are approximately linear. Therefore, creep
coefficient can be written as follows [11]:

ࢿࢉ ሺ࢚ି࢚૙ ሻ
‫׎‬ሺ࢚ǡ ࢚૙ ሻ ൌ Eq. (4. 2)
ࢿࢋ ሺ࢚૙ ሻ

Where; ߝ௖ ሺ‫ ݐ‬െ ‫ݐ‬଴ ሻ is the creep strain in duration of ሺ‫ ݐ‬െ ‫ݐ‬଴ ሻ and ߝ௘ ሺ‫ݐ‬଴ ሻ is the initial elastic
strain.

125
Chapter 4 – Time-dependent behaviour of concrete

1000 days 100 days


1 day t - t0 =1 min
Linear diagram boundary

Stress (σ)

Strain (ε-ε0)

Figure 4.4: Creep isochrones in concrete [11]

4.4. SHRINKAGE STRAIN

As the concrete shrinks, it compresses the steel reinforcement. The steel, in turn, imposes an
equal and opposite tensile force, ΔT on the concrete. This progressively increasing tensile force,
acting at some eccentricity to centroid of the concrete cross-section, produces elastic plus creep
strains and a resulting curvature of the section. The shrinkage-induced curvature (κsh) often
leads to significant load-independent deflection of the member. Magnitude of ΔT and hence the
shrinkage-induced curvature depends on the quantity and position of the reinforcement and the
size of the compressive concrete in the cross section, and hence on the extent of cracking. The
extent of cracking depends, of course, on the magnitude of the applied moment. Although
shrinkage strain is independent of stress, the shrinkage-induced curvature is not independent of
the external load. The effect of shrinkage-induced curvature on a previously cracked cross
section (κsh)cr is considerably greater than an uncracked cross-section (κsh)uncr as shown in Figure
(4.5) [12].

126
Chapter 4 – Time-dependent behaviour of concrete

Figure 4.5: Shrinkage-induced deformation and stresses in a singly reinforced beam [12]

4.5. EXISTING CREEP AND SHRINKAGE MODELS FOR LIGHTWEIGHT,


SELF-COMPACTING AND CONVENTIONAL CONCRETE

4.5.1. Creep and Shrinkage Models for Lightweight Concrete

4.5.1.1. Creep models for lightweight concrete

In the absence of testing to determine the creep strain and creep coefficient, there are several
models in the literature to use for design purposes. These models are developed by the
researchers and international codes, based on extensive experimental results. However, these
empirical models vary according to the application techniques, properties of the materials and
loading age as input data to ascertain the creep strain changes by time and related creep
coefficient.

Table (4.1) shows the existing models in codes of practice and some empirical models to
calculate the creep strain and coefficient in LWC. Since these models, especially the empirical
models have been developed for specific materials and testing setup, they may vary in accuracy
and complexity in calculations.
127
Chapter 4 – Time-dependent behaviour of concrete

Table 4.1: Existing creep models of lightweight concrete

Reference Creep model Notes


Best and Polivka (1959) ଵ εc : creep strain,
ߝ௖ ൌ ʹ͵ͳ‫ ݐ‬ൗସǤଵ
[14] t the time under load
ሺ‫ ݐ‬െ ‫ݐ‬଴ ሻ Ross hyperbolic expression
‫׎‬ሺ‫ݐ‬ǡ ‫ݐ‬଴ ሻ ൌ
‫ ܣ‬൅ ‫ܤ‬ሺ‫ ݐ‬െ ‫ݐ‬଴ ሻ A/B = 50 A=5.14E-2 ,

Mix A (x10-2) B (x10-2) A/B The polystyrene aggregate


A-0 2.9086 1.0346 28.1 concrete mixtures were produced
Ben Sabaa and by replacing 30, 50 and 70% of
A-30 4.2501 1.1101 38.3
Ravindrarajah (1997) the solid volume of the coarse
A-50 4.3306 0.9874 43.9
[15] aggregate with the polystyrene
A-70 5.212 1.2102 43.1
B-0 5.9703 1.1395 52.4 aggregate.
B-30 7.7086 1.2608 61.1
B-50 6.0769 1.3866 43.8
B-70 8.9621 1.2528 71.5
ሺ‫ ݐ‬െ ‫ݐ‬଴ ሻ (V/S) = 37.5 for a cylinder size of
‫׎‬ሺ‫ݐ‬ǡ ‫ݐ‬଴ ሻ ൌ ‫ߛ ߛ ߛ ׎‬
݂ ൅ ሺ‫ ݐ‬െ ‫ݐ‬଴ ሻ ௨ ோு ௏ௌ ௧బ 150 by 300 mm
௏ ‫׎‬௨ ൌ ʹǤ͵ͷ ultimate creep
ቄଵǤସଶൈଵ଴షమ ቀ ቁቅ
ACI-209R-92 (1997) ݂ ൌ ʹ͸݁ ௌ
coefficient
[16] (LWC and CC) ߛோு ൌ ͳǤʹ͹ െ ͲǤ͸͹ܴ‫ ݄ݎ݋݂ܪ‬൒ ͲǤͶ
௏ RH: relative humidity
ቀି଴Ǥ଴ଶଵଷ ቁ
ߛ௏ௌ ൌ ሺͳ ൅ ͳǤͳ͵݁ ௌ ሻ
ି଴Ǥଵଵ଼
ߛ௧బ ൌ ͳǤʹͷ‫ݐ‬଴
‫׎‬ሺ‫ݐ‬ǡ ‫ݐ‬଴ ሻ ൌ ‫׎‬଴ ߚ௖ ሺ‫ݐ‬ǡ ‫ݐ‬଴ ሻ T: temperature
‫׎‬଴ ൌ ‫׎‬ோுǡ் ߚሺ݂௖௠ିଶ଼ ሻߚሺ‫ݐ‬଴ ሻ ݂௖௠ିଶ଼ : mean compressive
଴Ǥଷ strength
‫ ݐ‬െ ‫ݐ‬଴
ߚ଴ ሺ‫ ݐ‬െ ‫ݐ‬଴ ሻ ൌ ቈ ቉ ‫׎‬଴ ǣ ݊‫ݐ݂݂݊݁݅ܿ݅݁݋ܿ݌݁݁ݎ݈݈ܿܽ݊݋݅ݐ݋‬
ߚுǡ் ൅ ሺ‫ ݐ‬െ ‫ݐ‬଴ ሻ
‫׎‬ோுǡ் ൌ ‫ ் ׎‬൅ ሾ‫׎‬ோு െ ͳሿ‫׎‬ଵǤଶ

‫ ் ׎‬ൌ ݁ ሾ଴Ǥ଴ଵହሺ்ିଶ଴ሻሿ
ߚுǡ் ൌ ߚு ߚ்
ͳ െ ܴ‫ܪ‬
‫׎‬ோு ൌ ሾͳ ൅ ߙଵ ሿߙଶ

ܸ
CEB MC-99 (1999) [17] ඩͲǤͳ ቎ ܵ ቏
ͷͲ

ͷǤ͵
ߚሺ݂௖௠ିଶ଼ ሻ ൌ
ඥ݂௖௠ିଶ଼ ȀͳͲ
ͳ
ߚሺ‫ݐ‬଴ ሻ ൌ
ͲǤͳ ൅ ‫ݐ‬଴଴Ǥଶ
ܸ
ߚு ൌ ͳͷͲሾͳ ൅ ሺͳǤʹ݄ሻଵ଼ ሿ ൬ ൰ ൅ ʹͷͲߙଷ ൑ ͳͷͲͲߙଷ
ܵ
ଵହ଴଴
ቂ ିହǤଵଶቃ
ߚ் ൌ ݁ ଶ଻ଷା்

ଷହ ଴Ǥ଻ ଷହ ଴Ǥଶ ଷହ ଴Ǥହ


ߙଵ ൌ ቂ ቃ , ߙଶ ൌ ቂ ቃ , ߙଵ ൌ ቂ ቃ
௙೎೘షమఴ ௙೎೘షమఴ ௙೎೘షమఴ
Similar to CEB MC-99 model except:
CEB-MC90 (1993)[18]
ߙଵ ൌ ߙଶ ൌ ߙଷ ൌ ͳ

128
Chapter 4 – Time-dependent behaviour of concrete

Table 4.1: Existing creep models of lightweight concrete (continued)

Reference Creep model Notes


‫׎‬ሺ‫ݐ‬ǡ ‫ݐ‬଴ ሻ ൌ ‫׎‬଴ ߚ௖ ሺ‫ݐ‬ǡ ‫ݐ‬଴ ሻ
‫׎‬଴ ൌ ‫׎‬ோு ߚሺ݂௖௠ିଶ଼ ሻߚሺ‫ݐ‬଴ ሻ
଴Ǥଷ
‫ ݐ‬െ ‫ݐ‬଴
ߚ଴ ሺ‫ ݐ‬െ ‫ݐ‬଴ ሻ ൌ ൤ ൨
ߚு ൅ ሺ‫ ݐ‬െ ‫ݐ‬଴ ሻ
ͳെ݄
‫׎‬ோு ൌ ሾͳ ൅ ߙଵ ሿߙଶ

ܸ
ඩͲǤͳ ቎ ܵ ቏
ͷͲ
EC-2 (2002) [19] tc is age of concrete where moist-
curing ends in days
ͷǤ͵
ߚሺ݂௖௠ିଶ଼ ሻ ൌ
ඥ݂௖௠ିଶ଼ ȀͳͲ
ͳ
ߚሺ‫ݐ‬଴ ሻ ൌ
ͲǤͳ ൅ ‫ݐ‬଴଴Ǥଶ
ܸ
ߚு ൌ ͳͷͲሾͳ ൅ ሺͳǤʹ݄ሻଵ଼ ሿ ൬ ൰ ൅ ʹͷͲߙଷ ൑ ͳͷͲͲߙଷ
ܵ
ଷହ ଴Ǥ଻ ଷହ ଴Ǥଶ ଷହ ଴Ǥହ
ߙଵ ൌ ቂ ቃ , ߙଶ ൌ ቂ ቃ , ߙଵ ൌ ቂ ቃ
௙೎೘షమఴ ௙೎೘షమఴ ௙೎೘షమఴ
‫׎‬ሺ‫ݐ‬ǡ ‫ݐ‬଴ ሻ
ʹሺ‫ ݐ‬െ ‫ݐ‬଴ ሻ଴Ǥଷ
ൌ ‫׎‬ሺ‫ݐ‬௖ ሻ ൥
ሺ‫ ݐ‬െ ‫ݐ‬଴ ሻ଴Ǥଷ ൅ ͳͶ
଴Ǥହ
͹ ଴Ǥହ ሺ‫ ݐ‬െ ‫ݐ‬଴ ሻ First two terms calculates the
GL2000 (Gardner) ൅൬ ൰ ቆ ቇ ൅ ʹǤͷሺͳ
‫ݐ‬଴ ሺ‫ݐ‬ െ ‫ݐ‬଴ ሻ ൅ ͹ basic creep and the third term is
model (2004) [20] ଴Ǥହ
ሺ‫ ݐ‬െ ‫ݐ‬଴ ሻ for drying creep
െ ͳǤͲͺ͸݄ଶ ሻ ቆ ቇ ൩
ሺ‫ ݐ‬െ ‫ݐ‬଴ ሻ ൅ ͹͹ሺܸȀܵሻଶ
଴Ǥହ ଴Ǥହ
ሺ‫ ݐ‬െ ‫ݐ‬଴ ሻ
‫׎‬ሺ‫ݐ‬௖ ሻ ൌ ൥ͳ െ ቆ ቇ ൩
ሺ‫ ݐ‬െ ‫ݐ‬଴ ሻ ൅ ͲǤͳʹሺܸȀܵሻଶ
‫׎‬ሺ‫ݐ‬ǡ ‫ݐ‬଴ ሻ ൌ ‫׎‬௕௖ ሺ‫ݐ‬ǡ ‫ݐ‬଴ ሻ ൅ ‫׎‬ௗ௖ ሺ‫ݐ‬ǡ ‫ݐ‬଴ ሻ
‫׎‬௕௖ ሺ‫ݐ‬ǡ ‫ݐ‬଴ ሻ ൌ ߚ௕௖ ሺ݂௖௠ ሻǤ ߚ௕௖ ሺ‫ݐ‬ǡ ‫ݐ‬଴ ሻ
ͳǤͺ
ߚ௕௖ ሺ݂௖௠ ሻ ൌ
ሺ݂௖௠ ሻ଴Ǥ଻

͵Ͳ
ߚ௕௖ ሺ‫ݐ‬ǡ ‫ݐ‬଴ ሻ ൌ ݈݊ ൭ቆ ൅ ͲǤͲ͵ͷቇ Ǥ ሺ‫ ݐ‬െ ‫ݐ‬଴ ሻ ൅ ͳ൱
‫ݐ‬଴Ǥ௔ௗ௝
‫׎‬ௗ௖ ሺ‫ݐ‬ǡ ‫ݐ‬଴ ሻ ൌ ߚௗ௖ ሺ݂௖௠ ሻǤ ߚሺܴǤ ‫ܪ‬ሻߚௗ௖ ሺ‫ݐ‬଴ ሻǤ ߚௗ௖ ሺ‫ݐ‬ǡ ‫ݐ‬଴ ሻ
Ͷͳʹ
ߚௗ௖ ሺ݂௖௠ ሻ ൌ
ሺ݂௖௠ ሻଵǤସ
FIB-2010 (2013) [21]
ͳ െ ܴ‫ܪ‬ൗͳͲͲ
ߚோǤு ൌ య
ඥͲǤͳ݄ȀͳͲͲ
ͳ
ߚௗ௖ ሺ‫ݐ‬଴ ሻ ൌ
ͲǤͳ ൅ ‫ݐ‬଴Ǥ௔ௗ௝ ଶ
‫ ݐ‬െ ‫ݐ‬଴ ఊ೟బ
ߚௗ௖ ሺ‫ݐ‬ǡ ‫ݐ‬଴ ሻ ൌ ൤ ൨
ߚ௛ ൅ ሺ‫ ݐ‬െ ‫ݐ‬଴ ሻ
ͳ
ߛ௧బ ൌ
͵Ǥͷ
ʹǤ͵ ൅
ඥ‫ݐ‬଴Ǥ௔ௗ௝
ߚ௛ ൌ ͳǤͷ݄ ൅ ʹͷͲߙ௖Ǥ௩

129
Chapter 4 – Time-dependent behaviour of concrete

4.5.1.2. Shrinkage models of lightweight concrete

Table (4.2) shows the existing models in codes of practice and some empirical models to
predict the shrinkage strain for LWC. Since these models, especially the empirical models have
been developed for specific materials and testing setup, they may vary in accuracy and
complexity in calculations.

Table 4.2: Existing models to predict shrinkage strain in lightweight concrete

Reference Shrinkage model Notes


Best and Polivka (1959) [14] ߝ௦௛ ൌ ʹ͵ͳ‫ ݐ‬ଵȀସǤଵ
ɂୱ୦ ൌ ɂୡୱୣ ൅ ɂୡୱୢ
‫כ‬
ɂୡୱୣ ൌ ߝ௖௦௘ ൈ ሺͳǤͲ െ ݁ ି଴Ǥଵ௧ ሻ
‫כ‬
ߝ௖௦௘ ൌ ሺͲǤͲ͸݂௖ᇱ െ ͳǤͲሻ ൈ ͷͲ ൈ ͳͲି଺
AS-3600 (2009) [22] ɂୡୱୢ ൌ ݇ଵ ݇ସ ɂୡୱୢǤୠ
‫כ‬
ɂୡୱୢǤୠ ൌ ሺͳǤͲ െ ͲǤͲͲͺ݂௖ᇱ ሻ ൈ ߝ௖௦ௗǤ௕
‫כ‬
ߝ௖௦ௗǤ௕ ൌ ሺͺͲͲǡ ͻͲͲǡ ͳͲͲͲሻ ൈ ͳͲି଺
For Sydney, Brisbane and Melbourne respectively
ሺ– െ ‫ݐ‬଴ ሻఈ
ɂୱ୦ ൌ ɂ
݂ ൅ ሺ– െ ‫ݐ‬଴ ሻఈ ୱ୦Ǥ୳
ɂ௦௛Ǥ௨ ൌ ͹ͺͲ ൈ ͳͲି଺ ߛ௦௛
షమ ሺ௏Ȁௌሻሽ
݂ ൌ ʹ͸ǤͲ݁ ሼଵǤସଶൈଵ଴
ߛ௦௛ ൌ ߛ௦௛ǡ௧௖ ߛ௦௛ǡோு ߛ௦௛ǡ௏ௌ ߛ௦௛ǡௗ ߛ௦௛ǡஏ ߛ௦௛ǡ௖ ߛ௦௛ǡ௔
ߛ௦௛ǡ௧௖ ൌ ͳǤʹͲʹ െ ͲǤʹ͵͵͹Ž‘‰ሺ‫ݐ‬଴ ሻ
ߛ௦௛ǡோு ൌ ͳǤͶ െ ͳǤͲʹ݄݂‫Ͳݎ݋‬ǤͶͲ ൑ ݄ ൑ ͲǤͺ
ACI-209R-92 (2008) [13](LWC and CC)
ߛ௦௛ǡோு ൌ ͵ǤͲ െ ͵ǤͲ݄݂‫Ͳݎ݋‬ǤͺͲ ൑ ݄ ൑ ͳ
ߛ௦௛ǡ௏ௌ ൌ ͳǤʹ݁ ሼିǤ଴଴ସ଻ଶሺ௏Ȁௌሻሽ
ߛ௦௛ǡஏ ൌ ͲǤ͵ ൅ ͲǤͲͳͶȲ
ߛ௦௛ǡ௖ ൌ ͲǤ͹ͷ ൅ ͲǤͲͲͲ͸ͳܿ ൒ ͳ
ߛ௦௛ǡ௔ ൌ ͲǤͻͷ ൅ ͲǤͲͲͺܽ ൒ ͳ
ߛ௦௛ǡௗ ൌ ͳǤʹ͵ െ ͲǤͲͲͳͷ݂݀‫ ݐݎ݋‬െ ‫ݐ‬଴ ൑ ͳ‫ݎܽ݁ݕ‬
ߛ௦௛ǡௗ ൌ ͳǤͳ͹ െ ͲǤͲͲͳͳͶ݂݀‫ ݐݎ݋‬െ ‫ݐ‬଴ ൐ ͳ‫ݎܽ݁ݕ‬

4.5.2. Creep and Shrinkage Models of Conventional Concrete

4.5.2.1. Creep models for conventional concrete

Table (4.3) shows the existing models in codes of practice and some empirical models to
estimate the creep coefficient in conventional concrete. Since these models especially the
empirical models, have been developed for specific materials and testing systems, they may
vary in accuracy and complexity in calculations.
130
Chapter 4 – Time-dependent behaviour of concrete

Table 4.3: Existing creep models of conventional concrete

Reference Creep model Notes


ሺ‫ ݐ‬െ ‫ݐ‬଴ ሻ
‫׎‬ሺ‫ݐ‬ǡ ‫ݐ‬଴ ሻ ൌ ‫ߛ ߛ ߛ ׎‬
݂ ൅ ሺ‫ ݐ‬െ ‫ݐ‬଴ ሻ ௨ ோு ௏ௌ ௧బ
௏ (V/S) = 37.5 for a cylinder size of 150 by
ACI 209R-92 ቄଵǤସଶൈଵ଴షమ ቀ ቁቅ
݂ ൌ ʹ͸݁ ௌ 300 mm
(1997) [16]
ߛோு ൌ ͳǤʹ͹ െ ͲǤ͸͹݄݂‫ ݄ݎ݋‬൒ ͲǤͶ ‫׎‬௨ ൌ ʹǤ͵ͷ ultimate creep coefficient
(LWC and CC) ௏
ቀି଴Ǥ଴ଶଵଷ ቁ h:relative humidity
ߛ௏ௌ ൌ ሺͳ ൅ ͳǤͳ͵݁ ௌ ሻ

ߛ௧బ ൌ ͳǤʹͷ‫ݐ‬଴ି଴Ǥଵଵ଼
‫׎‬ሺ‫ݐ‬ǡ ‫ݐ‬଴ ሻ ൌ ‫׎‬଴ ߚ௖ ሺ‫ݐ‬ǡ ‫ݐ‬଴ ሻ
‫׎‬଴ ൌ ‫׎‬ோுǡ் ߚሺ݂௖௠ିଶ଼ ሻߚሺ‫ݐ‬଴ ሻ
଴Ǥଷ
‫ ݐ‬െ ‫ݐ‬଴
ߚ଴ ሺ‫ ݐ‬െ ‫ݐ‬଴ ሻ ൌ ቈ ቉
ߚுǡ் ൅ ሺ‫ ݐ‬െ ‫ݐ‬଴ ሻ
‫׎‬ோுǡ் ൌ ‫ ் ׎‬൅ ሾ‫׎‬ோு െ ͳሿ‫׎‬ଵǤଶ

‫ ் ׎‬ൌ ݁ ሾ଴Ǥ଴ଵହሺ்ିଶ଴ሻሿ
ߚுǡ் ൌ ߚு ߚ்
ͳെ݄
‫׎‬ோு ൌ ሾͳ ൅ ߙଵ ሿߙଶ

ܸ T: temperature
CEB MC-99 ඩͲǤͳ ቎ ܵ ቏ ݂௖௠ିଶ଼ : mean compressive strength
(1999) [17] ͷͲ
‫׎‬଴ ǣ ݊‫ݐ݂݂݊݁݅ܿ݅݁݋ܿ݌݁݁ݎ݈݈ܿܽ݊݋݅ݐ݋‬
ͷǤ͵
ߚሺ݂௖௠ିଶ଼ ሻ ൌ
ඥ݂௖௠ିଶ଼ ȀͳͲ
ͳ
ߚሺ‫ݐ‬଴ ሻ ൌ
ͲǤͳ ൅ ‫ݐ‬଴଴Ǥଶ
ܸ
ߚு ൌ ͳͷͲሾͳ ൅ ሺͳǤʹ݄ሻଵ଼ ሿ ൬ ൰ ൅ ʹͷͲߙଷ ൑ ͳͷͲͲߙଷ
ܵ
ଵହ଴଴
ቂ ିହǤଵଶቃ
ߚ் ൌ ݁ ଶ଻ଷା்

ଷହ ଴Ǥ଻ ଷହ ଴Ǥଶ ଷହ ଴Ǥହ


ߙଵ ൌ ቂ ቃ , ߙଶ ൌ ቂ ቃ , ߙଵ ൌ ቂ ቃ
௙೎೘షమఴ ௙೎೘షమఴ ௙೎೘షమఴ
CEB MC90 Similar to CEB MC-99 model except:
(1993)[18] ߙଵ ൌ ߙଶ ൌ ߙଷ ൌ ͳ
‫׎‬ሺ‫ݐ‬ǡ ‫ݐ‬଴ ሻ ൌ ‫׎‬଴ ߚ௖ ሺ‫ݐ‬ǡ ‫ݐ‬଴ ሻ
‫׎‬଴ ൌ ‫׎‬ோு ߚሺ݂௖௠ିଶ଼ ሻߚሺ‫ݐ‬଴ ሻ
଴Ǥଷ
‫ ݐ‬െ ‫ݐ‬଴
ߚ଴ ሺ‫ ݐ‬െ ‫ݐ‬଴ ሻ ൌ ൤ ൨
ߚு ൅ ሺ‫ ݐ‬െ ‫ݐ‬଴ ሻ
ͳെ݄
‫׎‬ோு ൌ ሾͳ ൅ ߙଵ ሿߙଶ

ܸ
EC2 (2002) ඩͲǤͳ ቎ ܵ ቏
ͷͲ
tc is age of concrete where moist-curing
[19]
ends in days
ͷǤ͵
ߚሺ݂௖௠ିଶ଼ ሻ ൌ
ඥ݂௖௠ିଶ଼ ȀͳͲ
ͳ
ߚሺ‫ݐ‬଴ ሻ ൌ
ͲǤͳ ൅ ‫ݐ‬଴଴Ǥଶ
ܸ
ߚு ൌ ͳͷͲሾͳ ൅ ሺͳǤʹ݄ሻଵ଼ ሿ ൬ ൰ ൅ ʹͷͲߙଷ ൑ ͳͷͲͲߙଷ
ܵ
ଷହ ଴Ǥ଻ ଷହ ଴Ǥଶ ଷହ ଴Ǥହ
ߙଵ ൌ ቂ ቃ , ߙଶ ൌ ቂ ቃ , ߙଵ ൌ ቂ ቃ
௙೎೘షమఴ ௙೎೘షమఴ ௙೎೘షమఴ

131
Chapter 4 – Time-dependent behaviour of concrete

Table 4.3: Existing creep models for conventional concrete (continued)

Reference Creep model in CC Notes


‫׎‬ሺ‫ݐ‬ǡ ‫ݐ‬଴ ሻ
ʹሺ‫ ݐ‬െ ‫ݐ‬଴ ሻ଴Ǥଷ
ൌ ‫׎‬ሺ‫ݐ‬௖ ሻ ൥
ሺ‫ ݐ‬െ ‫ݐ‬଴ ሻ଴Ǥଷ ൅ ͳͶ
଴Ǥହ
GL2000 ͹ ଴Ǥହ ሺ‫ ݐ‬െ ‫ݐ‬଴ ሻ
(Gardner) ൅൬ ൰ ቆ ቇ ൅ ʹǤͷሺͳ First two terms calculates the basic creep
‫ݐ‬଴ ሺ‫ ݐ‬െ ‫ݐ‬଴ ሻ ൅ ͹
model (2004) ଴Ǥହ and the third term is for drying creep
ሺ‫ ݐ‬െ ‫ݐ‬଴ ሻ
[20] െ ͳǤͲͺ͸݄ଶ ሻ ቆ ቇ ൩
ሺ‫ ݐ‬െ ‫ݐ‬଴ ሻ ൅ ͹͹ሺܸȀܵሻଶ
଴Ǥହ ଴Ǥହ
ሺ‫ ݐ‬െ ‫ݐ‬଴ ሻ
‫׎‬ሺ‫ݐ‬௖ ሻ ൌ ൥ͳ െ ቆ ቇ ൩
ሺ‫ ݐ‬െ ‫ݐ‬଴ ሻ ൅ ͲǤͳʹሺܸȀܵሻଶ
‫׎‬ሺ‫ݐ‬ǡ ‫ݐ‬଴ ሻ ൌ ‫׎‬௕௖ ሺ‫ݐ‬ǡ ‫ݐ‬଴ ሻ ൅ ‫׎‬ௗ௖ ሺ‫ݐ‬ǡ ‫ݐ‬଴ ሻ
‫׎‬௕௖ ሺ‫ݐ‬ǡ ‫ݐ‬଴ ሻ ൌ ߚ௕௖ ሺ݂௖௠ ሻǤ ߚ௕௖ ሺ‫ݐ‬ǡ ‫ݐ‬଴ ሻ
ͳǤͺ
ߚ௕௖ ሺ݂௖௠ ሻ ൌ
ሺ݂௖௠ ሻ଴Ǥ଻

͵Ͳ
ߚ௕௖ ሺ‫ݐ‬ǡ ‫ݐ‬଴ ሻ ൌ ݈݊ ൭ቆ ൅ ͲǤͲ͵ͷቇ Ǥ ሺ‫ ݐ‬െ ‫ݐ‬଴ ሻ ൅ ͳ൱
‫ݐ‬଴Ǥ௔ௗ௝
‫׎‬ௗ௖ ሺ‫ݐ‬ǡ ‫ݐ‬଴ ሻ ൌ ߚௗ௖ ሺ݂௖௠ ሻǤ ߚሺܴǤ ‫ܪ‬ሻߚௗ௖ ሺ‫ݐ‬଴ ሻǤ ߚௗ௖ ሺ‫ݐ‬ǡ ‫ݐ‬଴ ሻ
Ͷͳʹ
ߚௗ௖ ሺ݂௖௠ ሻ ൌ
FIB-2010 ሺ݂௖௠ ሻଵǤସ
(2013) [21] ͳ െ ܴ‫ܪ‬ൗͳͲͲ
ߚோǤு ൌ య
ඥͲǤͳ݄ȀͳͲͲ
ͳ
ߚௗ௖ ሺ‫ݐ‬଴ ሻ ൌ
ͲǤͳ ൅ ‫ݐ‬଴Ǥ௔ௗ௝ ଶ
‫ ݐ‬െ ‫ݐ‬଴ ఊ೟బ
ߚௗ௖ ሺ‫ݐ‬ǡ ‫ݐ‬଴ ሻ ൌ ൤ ൨
ߚ௛ ൅ ሺ‫ ݐ‬െ ‫ݐ‬଴ ሻ
ͳ
ߛ௧బ ൌ
͵Ǥͷ
ʹǤ͵ ൅
ඥ‫ݐ‬଴Ǥ௔ௗ௝
ߚ௛ ൌ ͳǤͷ݄ ൅ ʹͷͲߙ௖Ǥ௩
Cu : the ultimate creep coefficient
‫ܭ‬௛௖ : relative humidity factor
– ଴Ǥ଺
‫׎‬ሺ‫ݐ‬ǡ ‫ݐ‬଴ ሻ ൌ ‫׎‬ ‫ܭ‬ௗ௖ : minimum member thickness factor
ͳͲ ൅ – ଴Ǥ଺ ୳ ‫ܭ‬௦௖ :concrete consistency factor
ACI 435-R95
‫ ୳׎‬ൌ ʹǤ͵ͷߛ஼ோ
(2003) [23] ‫ܭ‬௙௖ : fine aggregate content factor
ߛ஼ோ ൌ ‫ܭ‬௛௖ ‫ܭ‬ௗ௖ ‫ܭ‬௦௖ ‫ܭ‬௙௖ ‫ܭ‬௔௖

‫ܭ‬௧௖బ ൌ ͳ ௖
‫ܭ‬௔௖ : air content factor
For standard conditions
‫ܭ‬௧௖బ : age of concrete at load applications
factor
‫׎‬ሺ‫ݐ‬ǡ ‫ݐ‬଴ ሻ ൌ ݇ଶ ݇ଷ ݇ସ ݇ହ ‫׎‬௖Ǥ௕
ߙଶ ‫ ݐ‬଴Ǥ଼
݇ଶ ൌ ଴Ǥ଼
‫ ݐ‬൅ ͲǤͳͷ‫ݐ‬௛
ߙଶ ൌ ͳǤͲ ൅ ͳǤͳʹ݁ ି଴Ǥ଴଴଼௧೓
ʹǤ͹
݇ଷ ൌ ݂‫ ߬ݎ݋‬൒ ͳ݀ܽ‫ݕ‬
ሾͳ ൅ ݈‫݃݋‬ሺ߬ሻሿ
AS-3600 (2900)
݇ସ ൌ ͲǤ͹ǡ ͲǤ͸ͷǡ ͲǤ͸ǡ ͲǤͷ
[22]
For arid, interior, temperate inland and tropical
envir.
݇ହ ൌ ͳ ‫݂  ׷‬௖ᇱ ൑ ͷͲ‫ܽܲܯ‬
݇ହ ൌ ሺʹǤͲ െ ߙଷ ሻ െ ͲǤͲʹሺͳǤͲ െ ߙଷ ሻ݂௖ᇱ
ߙଷ ൌ ͲǤ͹Ȁሺ݇ସ ߙଶ ሻ
ି଴Ǥ଻଻ଵ
‫׎‬௖௖Ǥ௕ ൌ ͷͲǤͲ͵͸݂௖ᇱ  ‫ Ͳʹ ׷‬൑ ݂௖ᇱ ൑ ͳͲͲ‫ܽܲܯ‬

132
Chapter 4 – Time-dependent behaviour of concrete

4.5.2.2. Shrinkage models for conventional concrete

Table (4.4) shows the existing models in codes of practice and some empirical models to
calculate the shrinkage strain in CC. Since these models especially the empirical models have
been developed for specific materials and testing setup, they may vary in accuracy and
complexity in calculations.

Table 4.4: Existing shrinkage models for conventional concrete

Shrinkage Shrinkage models in CC Notes


– Free shrinkage- after7days
ɂୱ୦ ൌ ൬ ൰ ሺɂୱ୦ ሻ୳
͵ͷ ൅ – moist curing
– Free shrinkage- after1-3days
ɂୱ୦ ൌ ൬ ൰ ሺɂୱ୦ ሻ୳
ͷͷ ൅ – moist curing
ACI-435-R95 (2003) [23] ሺɂୱ୦ ሻ୳ ƒš ൌ ͹ͺͲ ൈ ͳͲି଺ ɀୱ୦
 ୱୠ : coefficient for cement
ɀୱ୦ ൌ  ୱ୦  ୱୢ  ୱୱ  ୱ୤  ୱୟୡ  ୱୠ ൌͳ content

For standard conditions All other K are the same as


defined for creep
ሺ– െ ‫ݐ‬଴ ሻఈ
ɂୱ୦ ሺ–ǡ ‫ݐ‬଴ ሻ ൌ ɂ
ˆ ൅ ሺ– െ ‫ݐ‬଴ ሻఈ ୱ୦Ǥ୳
ି଺
ɂ௦௛Ǥ௨ ൌ ͹ͺͲ ൈ ͳͲ ߛ௦௛
షమ
݂ ൌ ʹ͸ǤͲ݁ ሼଵǤସଶൈଵ଴ ሺ௏Ȁௌሻሽ
ߛ௦௛ ൌ ߛ௦௛ǡ௧௖ ߛ௦௛ǡோு ߛ௦௛ǡ௏ௌ ߛ௦௛ǡௗ ߛ௦௛ǡஏ ߛ௦௛ǡ௖ ߛ௦௛ǡ௔
ߛ௦௛ǡ௧௖ ൌ ͳǤʹͲʹ െ ͲǤʹ͵͵͹Ž‘‰ሺ‫ݐ‬଴ ሻ
ACI-209R-92 (2008) [13] ߛ௦௛ǡோு ൌ ͳǤͶ െ ͳǤͲʹ݄݂‫Ͳݎ݋‬ǤͶͲ ൑ ݄ ൑ ͲǤͺ
(LWC and CC) ߛ௦௛ǡோு ൌ ͵ǤͲ െ ͵ǤͲ݄݂‫Ͳݎ݋‬ǤͺͲ ൑ ݄ ൑ ͳ
ߛ௦௛ǡ௏ௌ ൌ ͳǤʹ݁ ሼିǤ଴଴ସ଻ଶሺ௏Ȁௌሻሽ
ߛ௦௛ǡஏ ൌ ͲǤ͵ ൅ ͲǤͲͳͶȲ
ߛ௦௛ǡ௖ ൌ ͲǤ͹ͷ ൅ ͲǤͲͲͲ͸ͳܿ ൒ ͳ
ߛ௦௛ǡ௔ ൌ ͲǤͻͷ ൅ ͲǤͲͲͺܽ ൒ ͳ
ߛ௦௛ǡௗ ൌ ͳǤʹ͵ െ ͲǤͲͲͳͷ݂݀‫ ݐݎ݋‬െ ‫ݐ‬଴ ൑ ͳ‫ݎܽ݁ݕ‬
ߛ௦௛ǡௗ ൌ ͳǤͳ͹ െ ͲǤͲͲͳͳͶ݂݀‫ ݐݎ݋‬െ ‫ݐ‬଴ ൐ ͳ‫ݎܽ݁ݕ‬
ɂୱ୦ ൌ ɂୡୱୣ ൅ ɂୡୱୢ
‫כ‬
ɂୡୱୣ ൌ ߝ௖௦௘ ൈ ሺͳǤͲ െ ݁ ି଴Ǥଵ௧ ሻ
‫כ‬
ߝ௖௦௘ ൌ ሺͲǤͲ͸݂௖ᇱ െ ͳǤͲሻ ൈ ͷͲ ൈ ͳͲି଺
ɂୡୱୢ ൌ ݇ଵ ݇ସ ɂୡୱୢǤୠ
‫כ‬
ɂୡୱୢǤୠ ൌ ሺͳǤͲ െ ͲǤͲͲͺ݂௖ᇱ ሻ ൈ ߝ௖௦ௗǤ௕
‫כ‬
ߝ௖௦ௗǤ௕ ൌ ሺͺͲͲǡ ͻͲͲǡ ͳͲͲͲሻ ൈ ͳͲି଺
AS-3600 (2900) [22] For Sydney, Brisbane and Melbourne
ߙଵ ‫ ݐ‬଴Ǥ଼
݇ଵ ൌ ଴Ǥ଼
‫ ݐ‬൅ ͲǤͳͷ‫ݐ‬௛
ߙଵ ൌ ͲǤͺ ൅ ͳǤʹ݁ ି଴Ǥ଴଴ହ௧೓
݇ସ ൌ ͲǤ͹ǡ ͲǤ͸ͷǡ ͲǤ͸
For arid, interior and temperate inland
environment

133
Chapter 4 – Time-dependent behaviour of concrete

4.5.3. Creep and Shrinkage Models in Self-compacting Concrete

4.5.3.1. Creep models of self-compacting concrete

Table (4.5) shows the existing models in codes of practice and some empirical models to
estimate the creep coefficient in self-compacting (self-consolidating) concrete. Since these
models, especially the empirical models have been developed for specific materials and testing
systems, they may vary in accuracy and complexity in calculations.

Table 4.5: Existing creep models for self-compacting concrete

Reference Creep model Notes


ߝ௖௥ ሺ‫ݐ‬ǡ ‫ݐ‬଴ ሻ The model and
ߪ௖ ሺ‫ݐ‬଴ ሻ ሺͳ െ ܴ‫ ܪ‬Τܴ‫ܪ‬଴ ሻ ͷǤ͵ ͳ related
Poppe and ൌ ൈ ቈͳ ൅ ቉ൈ ൈ
‫ܧ‬௖௜ ͲǤͶ͸ሺ݄Ȁ݄଴ ሻଵȀଷ ሺ݂ Τ
ඥ ௖௠ ݂௖௠଴ ሻ ሺͲǤͳ ൅ ሺ‫ݐ‬଴ Τ‫ݐ‬ଵ ሻ ሻ
଴Ǥଶ coefficients are
De
଴Ǥଷ based on CEB-
Schutter
ሺ‫ ݐ‬െ ‫ݐ‬଴ ሻΤ‫ݐ‬ଵ ͳ FIP model
(2005)[24] ൈ൦ ൪ ൤ ൨
ଵ଼ ሺ‫ ݐ‬െ ‫ݐ‬଴ ሻ ͲǤͲͳ ൅ ͳǤ͵͹ሺܿȀ‫݌‬ሻ
൬ͳͷͲ ൬ͳ ൅ ቀͳǤʹ ܴ‫ܪ‬ൗܴ‫ ܪ‬ቁ ൰ ݄Τ݄଴ ൅ ʹͷͲ൰ ൅ ൗ‫ݐ‬
଴ ଵ
square and
cylinder
specimens
‫ ݐ‬଴Ǥ଻
‫׎‬ሺ‫ݐ‬ǡ ‫ݐ‬଴ ሻ ൌ ߶  ‫߶  ׷‬௨ ൌ ͳǤ͹ͷ loaded at 1 day
Larson ͳ͸ ൅ ‫ ݐ‬଴Ǥ଻ ௨ square
(2006)
specimens
[25] ‫ ݐ‬଴Ǥ଺
‫׎‬ሺ‫ݐ‬ǡ ‫ݐ‬଴ ሻ ൌ ߶  ‫߶  ׷‬௨ ൌ ʹǤͲ loaded at 28
ʹͶ ൅ ‫ ݐ‬଴Ǥ଺ ௨
days
(based on ACI-
209-R(97))
The model and
related
Cordoba
‫ݐ‬ఈ coefficients are
(2007) ‫׎‬ሺ‫ݐ‬ǡ ‫ݐ‬଴ ሻ ൌ ߶
݀ ൅ ‫ݐ‬ఈ ௨ based on ACI-
[26]
209-R(97)
model
ିଵ λ=1.80 , μ=0.9
ܿ
ߝ௖௥ ሺ–ǡ ‫ ݐ‬ᇱ ǡ ‫ݐ‬଴ ሻ ൌ ɐᇱ௖௣ ൈ ɂᇱ௖௥ ሾͳ െ ݁‫݌ݔ‬ሼെͲǤͲͻሺ‫ ݐ‬െ ‫ ݐ‬ᇱ ሻ଴Ǥହସ ሽሿ ൈ ൭ͲǤͲͳͷ ൅ ͳǤ͵ͷ ൬ ൰൱  , α=2.1
‫݌‬

experimental
ǣ ݂‫ ݎ݋‬൏ ͲǤ͸ͷ parameters

ିଵ
ܿ
Aslani and ߝ௖௥ ሺ–ǡ ‫ ݐ‬ᇱ ǡ ‫ݐ‬଴ ሻ ൌ ɐᇱ௖௣ ൈ ɂᇱ௖௥ ሾͳ െ ݁‫݌ݔ‬ሼെͲǤͲͻሺ‫ ݐ‬െ ‫ ݐ‬ᇱ ሻ଴Ǥହସ ሽሿ ൈ ൭ͲǤͲͳͷ ൅ ͳǤͲͷ ൬ ൰൱
‫݌‬
Maia ௖
(2013) ǣ݂‫ ݎ݋‬൒ ͲǤ͸ͷ

[27] ߤ ൅ ߣǤ ߪሺ‫ݐ‬ǡ ‫ݐ‬଴ ሻ௔
ɐᇱ௖௣ ൌ
ͳെߢ
ߪሺ‫ݐ‬଴ ሻ
ߪሺ‫ݐ‬ǡ ‫ݐ‬଴ ሻ ൌ
݂௖௠ ሺ‫ݐ‬ሻ

ʹͺ ௡
݂௖௠ ሺ‫ݐ‬ሻ ൌ ݂௖ǡଶ଼ Ǥ ݁‫ ݌ݔ‬ቈ‫ ݏ‬ᇱ ቆͳ െ ൬ ൰ ቇ቉
‫ݐ‬

134
Chapter 4 – Time-dependent behaviour of concrete

4.5.3.2. Shrinkage models of self-compacting concrete

Table (4.6) shows the existing models in codes of practice and some empirical models to
calculate the shrinkage strain in SCC. Since these models, especially the empirical models have
been developed for specific materials and testing setup, they may vary in accuracy and
complexity in calculations.

Table 4.6: Existing shrinkage models for self-compacting concrete

Reference Shrinkage model Notes


ɂୱ୦ ሺ–ǡ ‫ݐ‬଴ ሻ
ͳ͸Ͳ ݂௖௠
ൌ൤ ൅ ͳͲߚ௦௖ ሺͻ െ ሻ൨
ͳ െ ߙሺ‫ ݓ‬Τܿሻ ݂௖௠଴
ൈ ሾെͳǤͷͷሺͳ െ ሺܴ‫ ܪ‬Τܴ‫ܪ‬଴ ሻଷ ሻሿ

Poppe and De Schutter ሺ‫ ݐ‬െ ‫ݐ‬଴ ሻΤ‫ݐ‬ଵ Model and coefficients are
(2005) [24] ൈ቎ ቏ based on CEB-FIP(90)
ሺ‫ ݐ‬െ ‫ݐ‬଴ ሻ
ͶͷǤͷሺ݄Τ݄଴ ሻଶ ൅ ൗ‫ݐ‬

ܿ
ߙ ൌ ͶǤͳ ൬ ൰ െ ͳǤͺ
‫݌‬
ܿ
ߛ ൌ െʹǤͷ ൬ ൰ ൅ ʹǤ͸
‫݌‬
‫ݐ‬ Square specimens
ሺɂୱ୦ ሻ௧ ൌ ൈ ͷͷͲ ൈ ͳͲି଺ 
ʹͲ ൅ ‫ݐ‬
Larson (2006) [25]
‫ݐ‬
ሺɂୱ୦ ሻ௧ ൌ ൈ ͸ͲͲ ൈ ͳͲି଺  Cylinder specimens
ʹͲ ൅ ‫ݐ‬
‫ݐ‬ఈ Model and coefficients are
Cordoba (2007) [26] ሺɂୱ୦ ሻ௧ ൌ ሺɂ ሻ
݂ ൅ ‫ ݐ‬ఈ ୱ୦ ௨ based on ACI-209R(97)
‫ݐ‬
ɂୱ୦ ൌ െ݇௦ ݇௛ ൬ ൰ ሺͲǤͷ͸ ൈ ͳͲଷ ሻ ൈ ‫ܣ‬
ͷͷ ൅ ‫ݐ‬ Model and coefficients are
‫ݐ‬ based on AASHTO (2004)
Khayat and Long (2010) [28] ʹ͸݁ ଴Ǥ଴ଵସଶሺ௏Ȁௌሻ ൅ ‫ݐ‬ ͳͲ͸Ͷ െ ͵Ǥ͹ሺܸȀܵሻ
݇௦ ൌ ቎ ‫ݐ‬ ቏൤ ൨ A: Cement type factor
ͻʹ͵
Ͷͷ ൅ ‫ݐ‬ Steam-cured condition
‫ ܣ‬ൌ ͲǤͻͳͺǡ ͳǤͲ͸ͷ

ɂୱ୦ ሺ–ǡ ‫ݐ‬଴ ሻ ൌ ɂୱ୦ ൤ͳ െ ݁‫ ݌ݔ‬൜െͲǤͳሺ‫ݐ‬



ሺିଶǤସቀ ቁାଶǤଷሻ
െ ‫ݐ‬଴ ሻ ௣ ൠ൨ RH: relative humidity
w/c : water to cement ratio
Aslani and Maia (2013) [27] ோு
ɂୱ୦ ൌ ൤െͷͲ ൅ ͹ͺ ቄͳ െ ݁‫ ݌ݔ‬ቀ ቁቅ ൅
ଵ଴଴ V/S volume to surface ratio


͵ͺǤ͵ Ž ‫ ݓ‬െ ͲǤͻʹ݈݊ ቀ ቁ െ ͷ ቂ݈݊ ቀ
௏Ȁௌ
ቁቃ ൨ ൈ c/p: cement to powder ratio
௖ ଵ଴

ሺͳͲିହ ሻǣ݂‫ ݎ݋‬൏ ͲǤ͸ͷ

135
Chapter 4 – Time-dependent behaviour of concrete

4.6. SOME EXPERIMENTS ON CREEP AND SHRINKAGE OF


LIGHTWEIGHT CONCRETE IN THE LITERATURE

Best et al. (1959) [14] reported the creep data in a concrete made with expanded shale
aggregates. The monitoring period was 18 months and the compressive strength of concrete at
28 days was in the range of 20 to 35 MPa. They found that contrary to current opinion, creep in
lightweight concrete was essentially equal to or less than that in conventional concrete of
comparable strength. All 20.6 MPa lightweight concretes had a lower creep strain and total
strain (instantaneous plus shrinkage plus creep) than the corresponding conventional concretes.

The total strains of the 35 MPa LWC specimens were equal to or lower than the total strain of
the corresponding CC.

Drying shrinkage of the 35 MPa LWC was only 36% to 46 % of the drying shrinkage of the
CC. The instantaneous strains of LWC were considerably greater, particularly for the 20MPa
concrete. Also, the instantaneous strains of the 20 MPa LWC were about 42 % to 93 % higher
than the instantaneous strain of CC.

4.7. BOND TRANSFER LENGTH IN LIGHTWEIGHT CONCRETE

Due to presence of steel in the reinforced concrete member, consideration of bond between
the steel bar and concrete and their interaction is inevitable. The concrete-reinforcement bond
between two adjacent cracks carries a certain amount of tensile force normal to the cracked
plane contributes to the overall stiffness of the member. With the perfect bond, no slip occurs
between concrete and reinforcement, whereas with a poor bond, relative displacement (slip) can
occur [29].

When the tensile force carried by concrete is transferred to the steel bar, the steel stress at the
cracked section increases under restriction by bond forces developed along a certain length of
the steel bars called bond transfer length on either side of the crack [30]. There are just limited
studies in the short-term definition of bond transfer length; and long-term variation is almost
unknown. By increasing the cracks, the service loading gradually moves the bond transfer
region toward the unloaded end. From the resulting strain curves, a more or less bilinear
decreasing trend are seen, in which, the transition point happening at the limit of the initial
transfer region [31].

Beeby (1972) assumed a fixed transfer length (S0) in short-term loading to transfer the bond
stress through the concrete to steel. He investigated the mechanism of primary crack formation
to reach the final pattern in this regard [5], [32]. Considering the serviceability of structure,
136
Chapter 4 – Time-dependent behaviour of concrete

time-dependent bond transfer length significantly affects the crack width, crack distribution and
deflection of concrete structure. Even in the case of inadequate provided development/splice
length, bonding may determine the ultimate capacity of the reinforced concrete member [33].

Measurement of the local bond stress and local slip along a stressed steel bar in a tension
member is difficult to ascertain and very sensitive to experimental errors [34]. For example,
based on different relationships, tensile strength will differ because of differences in stress
distribution. A comparison is made in Figure (4.6) which also shows that the tensile strength
increases with increasing concrete strength, but not at the same rate [4]. As explained in Chapter
Two, the tensile strength due the splitting strength is denoted ft and the tensile strength due to
the flexural test is known as the modulus of rupture fr.

Figure 4.6: Direct and indirect tensile strength and modulus of rupture vs. compressive strength [4]

The actual bond stress distribution is shown in Figure (4.7). However, an average value
(τb-average) is typically used to describe the force transfer between the materials. The total force
transferred to concrete is equivalent to the difference in tensile forces between adjacent cracks.
By allowing a reduction in tensile force between the cracks, the contribution of concrete stiffens
the response of reinforcement. The concept is called “tension stiffening” and is significantly
influenced by the cracks spacing [35].

Solokov et al. (2010) [36] noted that for beams with average and high reinforcement ratios
(ρ>0.5%), accurate predictions of tension stiffening by all the methods yields an error (standard
deviation) from 8.8% to 10.3%. However, predictions for the lightly reinforced beams (ρ<0.5%)
were far less accurate. These inaccuracies are due to the increased effect of the tensile concrete
which is a highly dispersed value.

Beeby and Scott (2004) [30] proposed the Equation (4.3) to calculate the average strains ߝ௦ଶ
in a tension member based on bond transfer length S0 and crack spacing λ for short-term loading.

ߣ ൌ ܵ଴ ߝ௦ଶ  Eq. (4. 3)

137
Chapter 4 – Time-dependent behaviour of concrete

Where: εs2 is the average strain in the reinforcement at a crack (concrete carries no tension),
So is bond transfer length and λ is crack spacing.

Figure 4.7: Bond transfer of tension reinforcement in concrete

Beeby and Scott (2004) [30] mentioned that the stiffness of an axially reinforced tension
member is directly related to the number of cracks. They also found that the bond transfer
length is shown to be proportional to the cover, and the variation in reinforcement strain is
linear in the region where it is affected by a crack. This was in agreement with their previous
studies. The number and the extent of cracks are controlled by the size and placement of the
reinforcing steel.

Bizindavy et al. (1999) [31] concluded the existence of an approximated bilinear relationship
between the bond transfer length and the relative load level with a constant value in the range of
service load levels (loads lower than the initial cracking load) followed by a linearly increasing
function up to failure. In this regard, an experimental study in the University of Durham shows
approximately constant bond stress over time under sustained loading [37].

4.7.1. Time-dependent Bond Transfer Length

From the tension stiffening concept, concrete does not crack suddenly and completely, but
undergoes progressive micro-cracking (strain softening). Due to the bond between steel and
concrete, the intact concrete between adjacent primary cracks carries considerable tensile force
138
Chapter 4 – Time-dependent behaviour of concrete

[38]. This bond makes the stress-strain characteristics of tensile concrete in cracked reinforced
concrete significantly different from the stress-strain characteristics of plain concrete. Since
bond stresses arise from the changes in the steel force along the length, the effect of bond
becomes more pronounced at end anchorages of reinforcing bars and in the vicinity of the
cracks.

As shown in Figures (4.8 a, b), the schematic stress-strain diagram for steel and concrete in
reinforced concrete under uniaxial tension is divided into three regions; pre-cracking, crack
development stage, and post-cracking. Under sustained load, σc gradually reduces; due to
cracking and bond breakdown caused by drying shrinkage and, to a lesser extent, due to tensile
creep. Early shrinkage reduces the cracking load. Under sustained service loads shrinkage
makes additional primary cracks with time and causes a time-dependent decay of the steel-
concrete bond. This decay of the bond is a short-term effect which occurs particularly rapidly in
large diameter specimens [39].

Figure 4.8: a) Tension stiffening model derived from uniaxial tension b) the equivalent concrete stress-strain
relationship [40]

Nilson (1968) developed a finite element model for reinforced concrete, considering the
influence of reinforcement, progressive cracking, bond stress transfer and non-linear material
properties. He provided a significant improvement on the Ngo and Scordelis (1967) model.
Goto (1971) applied tension to the reinforcement in a series of uniaxially reinforced concrete
prisms and presented the internal crack mechanisms and determination of the schematic
deformation diagram shown in Figure (4.9).

139
Chapter 4 – Time-dependent behaviour of concrete

Figure 4.9: Schematic deformation of concrete in a tension specimen [41]

4.7.2. Time-dependent Effects of Creep and Shrinkage on Bond Mechanism

Creep and shrinkage over time influence the behaviour of reinforced concrete members,
especially in cracked members [5]. Similar results in cracked and uncracked sections are evident
in the experimental investigation performed by Divakar and Dilger (1988), emphasising the
need to account for shrinkage and creep effects in long-term study of RC members [4].

Creep and shrinkage effects on the behaviour of the cracked RC members has been
recognised from the beginning of the 2nd half of 20th century [36]. There are extensive analytical
and numerical works carried out in the last two decades on the long-term behaviour of
composite members. However, very limited benchmark experimental data is currently available
in literature, e.g. (Alsamsam 1991; Bradford and Gilbert 1991; Johnson 1987; Roll 1971;
Wright et al. 1992 and Jiansheng et al. 2010)[12].

Tension-stiffening relationships were coupled with shrinkage and accompanying creep


effects. In most cases, tension stiffening relationships were obtained from shrunk experimental
RC members [36]. Concrete shrinkage strain continues to increase with time at a decreasing
rate. Other internal events such as bond slip or crack development around the bar reduce the
tension stiffening. There is no reason to suppose that this reduction will occur at the same rate
as creep or shrinkage. Shrinkage induced stresses significantly drop after cracking and they
almost disappear with yielding of reinforcement [42].

Scanlon (1971) presented a model based on the modification of the tensile branch of concrete
stress–strain relationship. This basic model was later considered and modified in different
investigations. Lin and Scordelis (1975), Gilbert and Warner (1978), Cope et al. (1979),
Damjanic and Owen (1984), Bazant and Oh (1984), Carreira and Chu (1986), Massicotte et al.
(1990), Prakhya and Morley (1990), and Kaklauskas and Ghaboussi (2001) are some
140
Chapter 4 – Time-dependent behaviour of concrete

researchers who have modified and developed the model. They describe cracking by a smeared
model assuming an average concrete tensile stress in a direction perpendicular to cracks [18].

Since assuming a constant bond stress and tension stiffening over the time cannot be an
adequate model for reinforced concrete, there are considerable time-dependent changes in the
average tensile stress caused in concrete. These changes are caused by tensile creep, restrained
shrinkage, and the time-dependent breakdown of bond due to shrinkage-induced and cover-
controlled cracking between the primary cracks [43]. Webster and Brooker (2006) [44] found
that effective tensile strength of the concrete, can even be the degree of restraint of shrinkage
movements.

The time-dependent effects of creep, shrinkage and temperature variation also exceed the
nonlinear behaviour of the RC member due to bond-slip, aggregate interlock at a crack and
dowel action of the reinforcing steel crossing a crack [45].

4.7.3. Bond Behaviour After Cracking and Before Yielding of Steel

First cracking occurs at the weakest cross-section where the concrete tensile stress reaches
the lower characteristic value of the tensile strength. Therefore, the stress in the concrete at the
crack drops to zero. The bond characteristics of materials determine the position of subsequent
cracks relative to the first cracks [46]. The zero tensile stress of concrete at each crack is rising
with the distance from the crack due to the steel-concrete bond of a maximum value (less than
the tensile strength of concrete) of mid-way between the adjacent cracks. The next crack will
then not form within S0 (see Figure (4.10)) of the first crack as the stresses in the concrete are
lower within this limit than outside. Slip at the concrete-steel interface in the region of
significant bond stress (S0 on either side of the crack) causes the crack to open. When the bond
from steel to concrete can no longer transfer sufficient tensile force to form an additional crack
between two existing cracks, the final established cracking state is reached.

Scott and Gill (1971) studied seven tensile specimens with gauged reinforcing bars and found
a linear variation in steel strain adjacent to a crack location and linear relation between bond
stresses with the load. This linear variation of stress implies a constant bond stress, which
initially suggests some form of plastic behaviour [37].

Figure (4.10a) indicates the mechanism by which cracks form in a uniaxial reinforced
concrete prism under tension. Scott and Beeby (2004) presented the schematic variation of
strain and stress in the region of a crack, as shown in Figure (4.10b) [30].

141
Chapter 4 – Time-dependent behaviour of concrete

Figure 4.10: Mechanism of cracking formation in a uniaxially reinforced concrete prism under tension [30]

Schematic distribution of forces, strains, normal stresses and bond stresses along a cracked
reinforced concrete member are illustrated in Figure (4.11)

Figure 4.11: Schematic distribution of forces, strains, normal stresses, and bond stresses in cracked member
[47]

142
Chapter 4 – Time-dependent behaviour of concrete

4.7.4. Experimental Investigation on the Ratio of Long-term Bond Transfer


Length to Initial Bond Transfer Length in Conventional Concrete

As part of the long-term cracking study at the University of New South Wales (UNSW) by
Nejadi and Gilbert (2004) [48], a total of eight fully restrained slab specimens with four
different reinforcement layouts were monitored for up to 150 days. They measured the effect of
shrinkage on development of time-dependent direct tension cracking due to restrained
deformation. Details of bar arrangement and concrete cover for the specimens are shown in
Table (4.7). Figure (4.12) displays the plan and elevation of the test specimens under pure
tension. The slab specimens are effectively anchored at both ends to ensure the complete
restraining of specimens. At mid-span of each specimen, the section was locally reduced to
make sure that first cracking always occur at this location.

Figure 4.12: Details of specimen in long-term bond transfer length test [48]

To delay the onset of drying shrinkage, the specimens were kept in their moulds for three
days and moist-cured continually in lab conditions. All slabs were anchored to the support at
age 3 days. The drying process also commenced after the same period. The time-dependent
development of drying shrinkage strain is also measured on unrestrained specimens of similar
dimension and thickness to the slab specimens.

All tests of companion specimens were carried out in agreement with the relevant parts of the
Australian Standard AS-1012.13 [49] . Two different ready mix batches of concrete were used
for slabs through the test according to Table (4.8).

143
Chapter 4 – Time-dependent behaviour of concrete

Table 4.7: Details of test specimens (reinforcement and dimensions)

Slab No. of bars Bar dia.(mm) Ast (mm2) Cs (mm) Cb (mm) S (mm)
RS1-a 3 12 339 109 44 185
RS1-b 3 12 339 109 44 185
RS2-a 3 10 236 110 46 185
RS2-b 3 10 236 110 46 185
RS3-a 2 10 157 145 46 300
RS3-b 2 10 157 145 46 300
RS4-a 4 10 314 115 46 120
RS4-a 4 10 314 115 46 120

Table 4.8: Material properties of concrete batch 1 (B.1) and batch 2 (B.2)

Age (days) 3 7 14 21 28
Concrete batch B.1 B.2 B.1 B.2 B.1 B.2 B.1 B.2 B.1 B.2
Compressive strength
8.17 10.7 13.7 17.4 20.7 25 22.9 27.5 24.3 28.4
(MPa)
Flexural tensile strength
1.91 2.47 3.15 3.1 3.43 3.77 3.77 3.97 3.98 4.04
(MPa)
Indirect tensile strength
1.55 1.6 1.97 2.1
(MPa)
Modulus of elasticity
13240 16130 17130 18940 21080 21750 22150 22840 22810 23210
(MPa)

4.7.5. Analytical Model of Bond Transfer Length

In a fully restrained RC member as shown in Figure (4.13), as concrete shrinks, the


restraining force N(t) gradually increases until the first crack appears. The resulting force
reduces to Ncr, immediately after the first crack appears and the stress away from the crack is
less than the tensile strength of concrete, ft (t). On either side of the crack, the concrete shortens
elastically, allowing the crack to open to a width w.

The slight increase in crack width with distance from the steel bar can be obtained by
integrating the reduction in tensile elastic strain in the concrete (from that at the level of the bar)
over the half crack spacing on either side of the crack.[50]

144
Chapter 4 – Time-dependent behaviour of concrete

Figure 4.13: First cracking in a restrained direct tension member [51]

According to Figure (4.14), Gilbert (1992) [51] derived the following expressions in
Equations (4.4) to (4.11) for the restraining force Ncr, the concrete and steel stresses away from
the crack σc1 and σs1 respectively, and the steel stress at the crack σs2.

a) Average concrete stress just after first cracking

b) Steel stress just after first cracking

Figure 4.14: Variation of average stress in steel and concrete

௡ఘ௙೎೟ ஺೎೟
ܰ௖௥ ൌ ஼ ା௡ఘሺଵା஼
 Eq. (4. 4)
భ భሻ

ே೎ೝ ିఙೞభ ஺ೞ೟ ே೎ೝ ሺଵା஼భ ሻ


ߪ௖ଵ ൌ ൌ  Eq. (4. 5)
஺೎ ୅೎

బ ଶ௦
ߪ௦ଵ ൌ  െ ଷ௅ିଶ௦ ߪ௦ଶ ൌ െ‫ܥ‬ଵ ߪ௦ଶ  Eq. (4. 6)


ߪ௦ଶ ൌ  ஺೎ೝ  Eq. (4. 7)
ೞ೟

బ ଶ௦
‫ܥ‬ଵ ൌ   Eq. (4. 8)
ଷ௅ିଶ௦ బ

‫כ‬
ఙೞభ ‫ି כ‬ఙ‫כ‬
ఙೞమ ଶ
ೞభ
ாೞ
‫ܮ‬ ൅ ݉ ாೞ
ቀଷ ‫ݏ‬଴ ൅ ‫ݓ‬ቁ ൌ ο‫ݑ‬ Eq. (4. 9)

145
Chapter 4 – Time-dependent behaviour of concrete

‫כ‬ ିଶ௦ ௠ ‫כ‬ ଷο௨ா


ߪ௦ଵ ൌ  ଷ௅ିଶ௦బ ߪ௦ଶ ൅ ଷ௅ିଶ௦ ೞ௠ Eq. (4. 10)
బ௠ బ

‫כ‬ ேሺஶሻ
ߪ௦ଶ ൌ  ஺ೞ೟
 Eq. (4. 11)

The bond transfer length S0 as presented in Equation (4.12) is related to bar diameter and ratio
of reinforcement. This expression was proposed earlier by Faver et al. (1983) [52] for a member
that contained deformed bars or welded wire mesh. Base and Murray (1982) [29] used a similar
expression.

ௗ್
ܵ଴ ൌ  Eq. (4. 12)
ଵ଴ఘ

Where; db is bar diameter (mm) and ρ is tensile reinforcement ratio in the section.

Bamonte and Gambarova (2007) [53] compared the bond-slip in a plain and a deformed bar
and showed the difference as illustrated in Figure (4.15).

Figure 4.15: Bond-slip mechanisms in a) smooth bars; b) deformed bars [53]

The code alterations regarding tension stiffening proposed by Beeby, Scott and Jones (2005)
are supporting the effect of bar diameter and reinforcement ratio on bond transfer length [54].
The good bond between steel and concrete increases the stiffening effect and is more significant
for low reinforcement ratios than for higher ones, reported by Massicotte et al. (1990) in Figure.

146
Chapter 4 – Time-dependent behaviour of concrete

(4.16) [29]. Horizontal and vertical axes in Figure (4.16) are presenting the deflection (Δ) and
load (T), respectively.

Figure 5.16: Tension stiffening effect in strain-stress diagram respect to bond quality and steel ratio [29]

Sokolovas et al. (2010) [36] noted that the free shrinkage tension stiffening effect is strongly
related to the reinforcement ratio and is considerably more pronounced for beams with smaller
reinforcement ratios.

Figures (4.17 a, b and c) respectively show a portion of a fully restrained member under
direct tension when a) all the shrinkage has occurred, and b) the final pattern of cracks has been
established, and c) it illustrates the average stresses in the concrete and steel caused by
shrinkage [41].

Gilbert (1992) [51] derived an expression for the final average crack spacing Sr and width in
a fully restrained member, the final restraining force in a member, and the final concrete and
steel stresses by enforcing the requirements of compatibility and equilibrium. With these
derivations, the bond transfer length S0 was assumed to remain constant over time, and the
supports of the member were presumed to be immovable. However, these assumptions may
introduce significant error, because the results from this experimental program indicate the
shrinkage effects on the deterioration of the bond at the concrete-steel interface and a gradual
increase in S0 with time.

From the average steel strain in the region along the steel bars, the steel stress remote from
the cracks after the first crack (σs1) and after all shrinkage cracking (σ*s1) is calculated.
Similarly, from the measured steel strains in region 2 (measured by electric strain gauges) and
obtaining the average value for these strain gauges between the reinforcement bars, the steel

147
Chapter 4 – Time-dependent behaviour of concrete

stress in the vicinity of the first crack immediately after the first cracking of σ s2, and after all
shrinkage cracking σ*s2 is also calculated.

By using Equations (4.4) to (4.11), the bond transfer length immediately after the first cracks
S0, and, after all shrinkage cracking S*0, was obtained. The results and ratio of S*0/ S0 are
presented in Table (4.9). Batches B1 and B2 are related to index ‘a’ and ‘b’ respectively.

Table 4.9: Measured no-bond length after first and all shrinkage cracking

Measured bond transfer length (mm)


Slab RS1- a RS1- b RS2- a RS2- b RS3- a RS3- b RS4- a RS4- b
After first cracking (S0) 261 215 228 201 246 342 256 259
After 150 days shrinkage 366 290 290 292 323 454 332 319
*
ܵ଴‫ כ‬Ȁܵ଴ 1.4 1.35 1.27 1.45 1.31 1.33 1.29 1.23

a) Portion of a restrained member after all cracking

b) Average concrete stress after all shrinkage cracking

c)
Steel stress after all shrinkage cracking
d)
Figure 4.17: Final concrete and steel stresses after direct tension cracking [51]

From Table (4.9), the average ratio of the final bond transfer length ܵ଴‫ כ‬to the initial value after
first cracking S0 is 1.33. Therefore, the final bond transfer length for long-term calculations may
be expressed as Equation (4.13)

148
Chapter 4 – Time-dependent behaviour of concrete

ܵ଴‫ כ‬ൌ ͳǤ͵͵ܵ଴  Eq. (4. 13)

From the test results, the extent to which shrinkage cracking can be controlled depends on
limiting the desired crack width and the amount and distribution of bonded reinforcement across
the crack.

The final crack width, crack spacing and steel stress at the crack are dependent on the steel
area, or more precisely, the reinforcement ratio. An increase in the steel area reduces the final
crack width and with more cracks developing, reduces the crack spacing.

With an increase in the steel area, the loss of stiffness at first cracking is reduced and,
therefore the restraining force after cracking is greater, but stress in the steel decreases at each
crack. With a larger restraining force, the stress in the concrete away from a crack tends to be
higher and consequently further cracking is more likely

4.8. EXISTING MODELS OF BOND-SLIP FOR LIGHTWEIGHT,


CONVENTIONAL AND SELF-COMPACTING CONCRETE

4.8.1. Bond-Slip Models for Lightweight Concrete

Table (4.10) shows the existing models used to predict the relationship between bond strength
and slip between the reinforcing bars and surrounding light weight concrete.

Table 4.10: Existing bond-slip models of light-weight concrete in literature

Reference Model for bond-slip Notes

τmax: peak bond stress

ఛ೘ೌೣ ஻ఙ σ: confining axisymmetric


ൌ‫ܣ‬൅
௙೟ ௙೟
Descending ቐ ௦೘ ஽ఙ radial pressure
ൌ‫ܥ‬൅
ௗ್ ௙೟
Malvar et al. (2003) [55] ft: tensile strength
ሼ߬ ൌ ߬௠௔௫
sm: slip at peak bond stress
௦ ఈ
Ascending ߬ ൌ ߬௠௔௫ ቀ ቁ
௦೘
db: nominal bar diameter

α, sr: empirical coefficients

Mitchel and Marzouk (2007) [56] investigated the bond-slip behavior of High-Strength
Light-Weight Concrete (HSLWC) and compared with the previous work at the same concrete

149
Chapter 4 – Time-dependent behaviour of concrete

lab by Alavi-Fard and Marzouk (2004) [57] on High-Strength Normal-Weight Concrete


(HSNWC). The same specimen size, casting position and test setup was used for all HSLWC
and HSNWC specimens. The experimental program consisted pull-out testing of 72 specimens
using 25 and 35 mm deformed reinforcement embedded in the specimens. The compressive
strength of specimens was about 80 MPa. The bond-slip behaviour in 25 mm and 35 mm steel
bars in HSLWC are presented in Figure (4.18).

1.2
35 mm (No. 11 bar)
1
Normalized bond stress

25 mm (No. 8 bar)
0.8

0.6

0.4

0.2

0
0 0.2 0.4 0.6 0.8 1 1.2
Normalized slaip

Figure 4.18: Normalized bond stress-slip for 25 and 35 mm bars under compression [56]

In comparison with HSNWC, they observed a similar manner in the bond-slip behaviour of
HSLSC and HSNWC. However, the maximum bond stress value (τmax) for 25 mm and 35 mm
bars were 10% lower and 10% greater in LWHSC than those for HSNWC respectively. The
HSLWC also showed a brittle failure during the pull-out tests.

The final slip value associated with bond failure was approximately five times greater than
the slip at the point of τmax. However, the shape of the bond-slip curve in both types of concrete
specimens was similar. The decrease in bond stress in the descending part was 50% to 60% for
HSLWC compared to 30% to 40% for HSNWC.

Based on the experimental results on the lightweight concrete containing expanded


polystyrene beads, a new model is proposed and verified in Chapter Five of this study.

4.8.2. Bond-Slip Models in Conventional Concrete

Table (4.11) shows the existing models used to predict the relationship between bond strength
and slip between the reinforcing bars and surrounding conventional concrete.
150
Chapter 4 – Time-dependent behaviour of concrete

Table 4.11: Existing bond-slip models of conventional concrete in literature

Reference Bond-slip models of conventional concrete Notes

݀ଶ ‫ݏ‬ ‫߬݌‬
ଶ ൌ ሺͳ ൅ ݊ߩሻ p represents the perimeter
Gupta and Maestrini (1990) ݀ ‫ܣ ݔ‬௦ ‫ܧ‬௦
(circumference) of the steel
[58] ‫ܧ‬௦ ‫ܣ‬௦
݊ ൌ ǡ ߩൌ bar
‫ܧ‬௖ ‫ܣ‬௖
߬ ൌ ሺͶͳǤ͹ െ ͲǤʹ݂௦ ሻ‫ ݏ‬଴Ǥ଼
Plain round bar Where τ =bond stress, in
N/mm2 , x =distance from
߬ ൌ ሺͷͷ െ ͲǤͷ‫ݔ‬ሻ‫ ݏ‬଴Ǥ଼
Kankam(1997) [59] centre of embedment, in
Cold-worked ribbed bar
mms; and s =local slip of
߬ ൌ ሺ͵ͷ െ ͲǤ͵‫ݔ‬ሻ‫ ݏ‬଴Ǥହ
bar, in mms, fs=steel stress
Hot-rolled ribbed bar

Where X = distance from


loaded face; and ඥ݂௖ᇱ =
Nilson (1971) ߬ ൌ ͵ͳͲͲሺͳǤͶ͵‫ ݔ‬൅ ͳǤͷሻ‫ݏ‬ඥ݂௖ᇱ
measure of tensile strength
of concrete.
߬ ൌ ͳǤͻͷ ൈ ͳͲ଺ ‫ݏ‬ െ ʹǤ͵ͷ ൈ ͳͲଽ ‫ ݏ‬ଶ ൅ ͳǤ͵ͻ ൈ ͳͲଵଶ ‫ ݏ‬ଷ
Mirza and Houde (1979) [60] s =local slip of bar, in mm
െ ͲǤ͵͵ ൈ ͳͲଵହ ‫ ݏ‬ସ

where Φ , ψ , α are
Rehm (1961) [61] ߬ ൌ ݂௖Ǥ௖௨௕ ሺ߮‫ ݏ‬ఈ േ Ȳ௦ ሻ theoretical or experimental
constants

Nilson (1968) [62] ߬ ൌ ͻͻͺǤͶ‫ ݏ‬െ ͷͺͶͲͲ‫ ݏ‬ଶ ൅ ͺͷʹʹͲͲ‫ ݏ‬ଷ

Martin (1973) [63] ߬ ൌ ߬଴ ൅ ߙ‫ ݏ‬ఉ

Mirza and Houde (1979) [60] ߬ ൌ ͷ͵ͻǤͺ‫ ݏ‬െ ʹͷ͸ͳͲ‫ ݏ‬ଶ ൅ ͷͻʹʹͲͲ‫ ݏ‬ଷ െ ͷͷ͹ͶͲͲͲ‫ ݏ‬ସ
௦ ఈ
Eligehausen et al. (1983) [64] ߬ ൌ ߬௠௔௫ ቀ ቁ Ͳ ൑ ‫ ݏ‬൑ ‫ݏ‬ଵ s, s1 (Figure 4.19)
௦భ

߬ ൌ ߬௠௔௫  ‫ݏ  ׷‬ଵ ൑ ‫ ݏ‬൑ ‫ݏ‬ଶ


CEB-FIP (1993) [18]
‫ ݏ‬െ ‫ݏ‬ଶ
Huang et al. (1996) [65] ߬ ൌ ߬௠௔௫ െ ൫߬௠௔௫ െ ߬௙ ൯ ൬ ൰ǣ‫ݏ‬ଶ ൑ ‫ ݏ‬൑ ‫ݏ‬ଷ s, s1, s2 , s3 (Figure 4.19)
‫ݏ‬ଷ െ ‫ݏ‬ଶ
Harajli et al (1995) [66]
߬ ൌ ߬௙  ‫ݏ  ׷‬ଷ ൏ ‫ݏ‬

151
Chapter 4 – Time-dependent behaviour of concrete

Table 4.11: Existing bond-slip models of conventional concrete in literature (continued)

Reference Bond-slip models of conventional concrete Notes


߬௠௔௫
߬ ൌ ି஻Ǥ௟௡ሺ஻Τ஽ ሻȀሺ஻ି஽ሻ ሺ݁ ஻௦ ሻ
ሾ݁ െ ݁ ି஽Ǥ௟௡ሺ஻Τ஽ ሻȀሺ஻ି஽ሻ ሿ Kst =Ast/nSst db
െ ݁ ஽௦ confinement factor
ͲǤͲʹͷͶ ൅ ‫ܭ‬௦௧
‫ܤ‬ൌ  Kco =c/db combined
െͲǤͲʹ͵ʹ െ ͺǤ͵Ͷ‫ܭ‬௦௧
parameter of bar diameter
െͳǤͳ ൑ ‫ ܤ‬൑ െͲǤͳʹ
Wu and Zhao (2012)
ͲǤ͹͵ͳͷ ൅ ‫ܭ‬ and concrete
‫ ܦ‬ൌ ͵݈݊ ൬ െ ͲǤͳ͵൰ െ ͵Ǥ͵͹ͷ
ͷǤͳ͹͸ ൅ ͲǤ͵͵͵͵‫ܭ‬
Ksi : size effect factor
‫ ܭ‬ൌ ሺ‫ܭ‬௖௢ ൅ ‫ܭܨ‬௦௧ ሻ‫ܭ‬௦௜ 
Rr: ratio of the rib height to
݀௕ ଴ǤଶିଵǤ଼ோೝ
‫ܭ‬௦௜ ൌ ൬ ൰ the rib spacing
ʹͷǤͶ
Ͷ߬௠௔௫
߬ൌ ‫ ݏ ׷ ݏ‬൏ ͲǤͳ‫ݏ‬௣௘௔௞
ܵ௣௘௔௞

‫ ݏ‬െ ‫ݏ‬௣௘௔௞
߬ ൌ ߬௠௔௫ ൥ͳ െ ͲǤ͸ ቆ ቇ ൩ ǣͲǤͳ‫ݏ‬௣௘௔௞ ൑ ‫ݏ‬
ͲǤͻ‫ݏ‬௣௘௔௞

Murcia-Delso et al. (2011) ൑ ‫ݏ‬௣௘௔௞  For Speak and SR, see
[67] ߬ ൌ ߬௠௔௫  ‫ݏ  ׷‬௣௘௔௞ ൑ ‫ ݏ‬൑ ͳǤͳ‫ݏ‬௣௘௔௞  Figure (4.20)
‫ ݏ‬െ ͳǤͳ‫ݏ‬௣௘௔௞
߬ ൌ ߬௠௔௫ ቈͳ െ ͲǤ͹ͷ ቆ ቇ቉ǣͳǤͳ‫ݏ‬௣௘௔௞
‫ݏ‬ோ െ ͲǤͻ‫ݏ‬௣௘௔௞
൑ ‫ ݏ‬൑ ‫ݏ‬ோ 
߬ ൌ ͲǤʹͷ߬௠௔௫  ‫ ݏ ׷‬൒ ‫ݏ‬ோ
߬ ൌ ͳͻǤ͵͸‫ ݏ‬଴Ǥହଵ  ‫݂  ׷‬௖ᇱ ൏ ͷͲ‫ܽܲܯ‬
Barbosa (2001) [68]
߬ ൌ ͵ʹǤͷͺ‫ ݏ‬଴Ǥସ଼  ‫݂  ׷‬௖ᇱ ൒ ͷͲ‫ܽܲܯ‬

଴Ǥ଺ ‫ݏ‬ ‫ݏ‬௣ଵ : slip value at ߬௠௔௫
߬ ൌ ʹǤͷ݂௖ᇱ ቆ ቇ
‫ݏ‬௣ଵ
Oh and Kim (2007) [69] ߬௡ ൌ Ͳ‫ ݏ ׷‬൏ ‫ݏ‬௥ሺ௡ିଵሻ Α: empirical factor

‫ ݏ‬െ ‫ݏ‬௥ሺ௡ିଵሻ ߬௡ : bond stress for repeated


߬௡ ሺ‫ݏ‬ሻ ൌ ߬௠௔௫ ቆ ቇ‫ ݏ ׷‬൒ ‫ݏ‬௥ሺ௡ିଵሻ
‫ݏ‬௣ଵ loading

The parameters s, s1, s2 and s3 in the Equation presented by CEB-FIP (1993) are shown in
Figure (4.19). This model is approved later by Huang et al. (1996), and Harajli et al (1995).

152
Chapter 4 – Time-dependent behaviour of concrete

Figure 4.19: Bond-slip model in reinforced concrete (CEB-FIP, 1993)[18]

The parameters in the equation presented by Murcia-Delso et al. (2011) are shown in Figure
(4.20).

Figure 4.20: The parameters for analytical models developed by Murcia-Delso et al. (2011)[67]

4.8.3. Bond-Slip Models in Self-compacting Concrete

Table (4.12) shows the existing models of predicting the relationship between bond strength
and slip between the reinforcing bars and surrounding self-compacting concrete.

153
Chapter 4 – Time-dependent behaviour of concrete

Table 4.12: Existing bond-slip models for self-compacting concrete in literature

Reference Model Notes


Model based on Ceb-Fip 195/197 (1993) [18]

Non-Confined
Confined concrete
concrete
‫ݏ‬ଵ 1.0 mm 0.6 mm
‫ ݏ‬ఈ
߬ ൌ ߬௠௔௫ Ǥ ൬ ൰ ‫ݏ‬ଶ 3.0 mm 0.6 mm
‫ݏ‬ଵ
‫ݏ‬ଷ Ribs interval 1.0 mm
Ͳ ൑ ‫ ݏ‬൑ ‫ݏ‬ଵ
ߙ 0.4 0.4
߬ ൌ ߬௠௔௫
߬௠௔௫ ʹǤͷඥ݂௖ᇱ ʹǤͲඥ݂௖ᇱ
‫ݏ‬ଵ ൑ ‫ ݏ‬൑ ‫ݏ‬ଶ
Almeida Filho ‫ ݏ‬െ ‫ݏ‬ଶ ߬௨ ͲǤͶ߬௠௔௫ ͲǤͳͷ߬௠௔௫
et al. (2008) ߬ ൌ ߬௠௔௫ െ ሺ߬௠௔௫ െ ߬௨ ሻ ൬ ൰
‫ݏ‬ଷ െ ‫ݏ‬ଶ Model based on Huang et al. (1996) [65]
[70]
‫ݏ‬ଶ ൑ ‫ ݏ‬൑ ‫ݏ‬ଷ Confined Non-Confined
concrete concrete
߬ ൌ ߬௨ 
‫ݏ‬ଵ 0.5 mm 1.0 mm
‫ݏ‬ଷ ൏ ‫ݏ‬
‫ݏ‬ଶ 1.5 mm 3.0 mm
‫ݏ‬ଷ Ribs interval Ribs interval
ߙ 0.3 0.4
߬௠௔௫ ͲǤͶ݂௖௠ ͲǤͶ݂௖௠
߬௨ ͲǤͶ߬௠௔௫ ͲǤͳͷ߬௠௔௫

4.9. SUMMARY

Following the description of the mechanical properties of EPS lightweight concrete in


Chapter 3, the time-dependent properties are discussed in Chapter 4. A brief description of the
creep and shrinkage effects on the deflection behaviour of reinforced concrete structures over
time is given also. Other effective parameters in the moment-curvature diagram of reinforced
concrete members subjected to the simultaneous effects of the sustained loading and creep and
shrinkage are explained.

The existing models used to predict the creep and shrinkage behaviour of lightweight
concrete are presented. Also the existing models for conventional and self-compacting concrete
are given and compared with the models for lightweight concrete. The experimental
investigations and the existing models show a higher range of creep and shrinkage for
lightweight concrete compared to the conventional and self-compacting concrete.

Characteristics of the steel-concrete interface are discussed as an effective parameter both in


instantaneous and time-dependent deflection of reinforced concrete structures. Some existing
models to estimate the bond-slip behaviour and the maximum bond stress in lightweight,
conventional and self-compacting concrete are given as well.

154
Chapter 4 – Time-dependent behaviour of concrete

The effect of bond transfer length on the deflecting and cracking behaviour of reinforced
concrete structures are also investigated and the relative models in the literature are presented.

The results of an experimental study by Nejadi and Gilbert (2004) to measure the time-
dependent bond-transfer length in a series of conventional concrete slabs are broadly evaluated.

The bond-slip changes between the reinforcing bars and the surrounding concrete is among
the most important parameters to influence the time-dependent and instantaneous deflection
behaviour of concrete structures. The existing models for lightweight, conventional and self-
compacting concrete are presented and compared.

155
Chapter 4 – Time-dependent behaviour of concrete

REFERENCES

1. Gilbert, R.I., Time effects in concrete structures. Developments in civil engineering, 1988. 23.
2. Unnikrishna Pillai, S. and D. Menon, Reinforced Concrete Design Tata Mc Graw Hill, New
Delhi, India,, 2003.
3. Nejadi, S., Time-dependent cracking and crack control in reinforced concrete structures. 2005,
PhD thesis, The University of New South Wales Sydney, Australia.
4. Eigelaar, E.M., Deflections of reinforced concrete flat slabs. 2010, Stellenbosch: University of
Stellenbosch.
5. Nie, J. and C.S. Cai, Deflection of cracked RC beams under sustained loading. Journal of
Structural Engineering, 2000. 126(6): p. 708-716.
6. Reichard, T.W., Creep and drying shrinkage of lightweight and normal weight concretes.
National Bureau of Standards, Washington, D.C., Monograph 74, 1964.
7. Aslani, F., Experimental and numerical study of time-dependent behaviour of reinforced self-
compacting concrete slabs. PhD dissertation, University of Technology Sydney (UTS), 2014.
8. Bazant, Z.P., Criteria for rational prediction of creep and shrinkage of concrete. ACI SPECIAL
PUBLICATIONS, 2000. 194: p. 237-260.
9. Chia, K.-S., et al., Experimental study on creep and shrinkage of high-performance ultra-
lightweight cement composite of 60 MPa. Structural Engineering and Mechanics, 2014. 50(5): p.
635-652.
10. Zhang, Q., et al. Long-term creep properties of cementitious materials–comparing compression
tests on concrete with microindentation tests on cement. in American Society of Civil Engineers.
2013.
11. Bazant, Z.P. and R. L'Hermite, Mathematical modeling of creep and shrinkage of concrete.
1988: Wiley Chichester.
12. Gilbert, R.I., Creep and Shrinkage Induced Deflections in RC Beams and Slabs. ACI , Special
Publication, 2012. 284(13): p. 1-16.
13. ACI-209-2R, Guide for Modeling and Calculating Shrinkage and Creep in Hardened Concrete.
Reported by ACI Committee 209, 2008.
14. Best, C.H. and M. Polivka Creep of lightweight concrete. Magazine of Concrete Research, 1959.
11, 129-134.
15. Sabaa, B. and R.S. Ravindrarajah. Engineering properties of lightweight concrete containing
crushed expanded polystyrene waste. in Materials Research Society, 1997, Fall Meeting,
Symposium MM, Advances in Materials for Cementitious Composites December. 1997.
16. ACI-209-2R, Prediction of Creep, Shrinkage, and Temperature Effects in Concrete Structures.
Reported by ACI Committee 209, 1997.
17. CEB-99, Concrete, Structural, Textbook on Behaviour, Design and Performance. Updated
knowledge of the CEB/FIP Model Code 1990. fib Bulletin 2, V.2, Fédération Internationale du
Béton, Lausanne, Switzerland, 1999. 1: p. 37-52.
18. béton, C.e.-i.d., CEB-FIP model code 1990: design code. 1993: CEB Bulletin d’Information
No.213/214, Comité EuroInternational du Béton, Lausanne, Switzerland, Telford.
19. EC-2, E., Design of Concrete Structures, Part 1: General Rules and Rules for buildings.
Commission of European Communities ENV, German version EN 1992-1-1:2004 + AC: 2010,
The Concrete Centre: Blackwater, Camberley, UK, . 2004: p. 1-1.
20. Gardner, N., Comparison of prediction provisions for drying shrinkage and creep of normal-
strength concretes. Canadian Journal of Civil Engineering, 2004. 31(5): p. 767-775.
21. du Béton, F.I., fib Model Code for Concrete Structures 2010. 2013, Lausanne, fib - federation
internationale du beton, John Wiley & Sons, 1 Oct 2013 - Technology & Engineering - 500
pages.
156
Chapter 4 – Time-dependent behaviour of concrete

22. AS-3600-09, Concrete structures. Australian Standard, 2009.


23. ACI-435-R95, Control of Deflection in Concrete Structures. American Concrete Institute (ACI)
Committee 435, 2003.
24. Poppe, A.-M. and G. De Schutter. Creep and shrinkage of self-compacting concrete. in First
International Symposium on Design, Performance and Use of Self-Consolidating Concrete,
China. 2005.
25. Larson, K.H., Evaluating the time-dependent deformations and bond characteristics of a self-
consolidating concrete mix and the implication for pretensioned bridge applications. 2006,
Kansas State University.
26. Cordoba, B., Creep and shrinkage of self-consolidating concrete (SCC). 2007: University of
Wyoming.
27. Aslani, F. and L. Maia, Creep and shrinkage of high-strength self-compacting concrete:
experimental and analytical analysis. Magazine of Concrete Research, 2013. 65(17): p. 1044-
1058.
28. Khayat, K.H. and W.J. Long, Shrinkage of precast, prestressed self-consolidating concrete. ACI
Materials Journal, 2010. 107(3).
29. Massicotte, B., A.E. Elwi, and J.G. MacGregor, Tension-stiffening model for planar reinforced
concrete members. Journal of Structural Engineering, 1990. 116(11): p. 3039-3058.
30. Beeby, A. and R. Scott, Cracking and deformation of axially reinforced members subjected to
pure tension. Magazine of concrete research., 2005. 57(10): p. 611-621.
31. Bizindavyi, L. and K. Neale, Transfer lengths and bond strengths for composites bonded to
concrete. Journal of composites for construction, 1999. 3(4): p. 153-160.
32. Alih, S. and A. Khelil. Tension Stiffening Parameter in Composite Concrete Reinforced with
Inoxydable Steel: Laboratory and Finite Element Analysis. in Proceedings of World Academy of
Science, Engineering and Technology. 2012. World Academy of Science, Engineering and
Technology.
33. Wang, H., An analytical study of bond strength associated with splitting of concrete cover.
Engineering Structures, 2009. 31(4): p. 968-975.
34. Chan, H., Y. Cheung, and Y. Huang, Crack analysis of reinforced concrete tension members.
Journal of Structural Engineering, 1992. 118(8): p. 2118-2132.
35. Vogel, H.M., Serviceability of concrete beams reinforced with FRP and concrete prisms
prestressed with CFRP. University of Manitoba (Canada), 2011. 2011.NR78922. .
36. Sokolovas, A., et al., Tension-Stiffening Model Based on Test Data of RC Beams. Modern
building materials, structures and techniques, 2010. Vilnius, Lithuania, The 10th international
Conference: p. 5., 2010.
37. Wenkenbach, I., Tension Stiffening in Reinforced Concrete Members with Large Diameter
Reinforcement. Durham University, undergraduate thesis. Available at Durham E-Theses
Online:< http://etheses. dur. ac. uk/3250/>[Accessed 15/08/2012], 2011.
38. Behfarnia, K., The effect of tension stiffening on the behaviour of R/C beams. Asian Journal of
Civil Engineering (Building and Housing), 2009. 10(3): p. 243-255.
39. Vakhshouri;, B. and S. Nejadi, Time-dependent bond transfer length under pure tension in one
way slabs. Structural Engineering and Mechanics, 2016. 60(2).
40. Ian Gilbert, R., Tension Stiffening in Lightly Reinforced Concrete Slabs. Journal of Structural
Engineering, 2007. 133(6): p. 899-903.
41. Vakhshouri, B. and S. Nejadi, Experimentla study of time-dependant bond-transfer length under
pure tension in slabs. IOS press, Construction materilas and structures, proceeding of the 1st int.
conferenec on construction materials and structures 2014: p. 1044-1051.
42. Kaklauskas, G. and J. Ghaboussi, Stress-strain relations for cracked tensile concrete from RC
beam tests. Journal of Structural Engineering, 2001. 127(1): p. 64-73.

157
Chapter 4 – Time-dependent behaviour of concrete

43. Gilbert, R.I., Closure to “Tension Stiffening in Lightly Reinforced Concrete Slabs” by R. Ian
Gilbert. Journal of structural engineering, 2008. 134(7): p. 1264-1265.
44. Bond, A., et al., How to design concrete structures using Eurocode 2. 2006: Concrete Centre.
45. Kwak, H.-G. and F.C. Filippou, Finite element analysis of reinforced concrete structures under
monotonic loads. 1990: Department of Civil Engineering, University of California Berkeley,
CA, USA.
46. Mihai, P., I. Hirhui, and B. Rosca, Reinforcement using the Finite Element Method. Journal of
Applied Sciences, 2010. 10(9): p. 738-744.
47. Vakhshouri, B. and S. Nejadi. Experimental study of time dependent bond transfer length under
pure tension in slabs. in International Conference on Construction Materials and Structures.
2014. IOS Press BV.
48. Nejadi, S. and R.I. Gilbert, Shrinkage cracking and crack control in restrained reinforced
concrete members. ACI Structural Journal, 2004. 101(6).
49. AS-1012.13, Methods of testing concrete, method 13-determination of the drying shrinkage of
concrete for samples prepared in the field or in the laboratory. Australian standard, 2014.
50. Wu, H.Q. and R.I. Gilbert, Modeling short-term tension stiffening in reinforced concrete prisms
using a continuum-based finite element model. Engineering Structures, 2009. 31(10): p. 2380-
2391.
51. Gilbert, R.I., Shrinkage cracking in fully restrained concrete members. ACI Structural Journal,
1992. 89(2).
52. Faver, R., M. Koprna, and A. Radojicic, Effects differes, fissuration et deformations des
structures en beton. Manual du Euro-International Du beton, Ecole Polytechnique Federale de
Lausanne, Editions Georgi, Saint-Saphorin, Suisse, 1983.
53. Bamonte, P. and P. Gambarova, High-bond bars in NSC and HPC: study on size effect and on
the local bond stress-slip law. Journal of Structural Engineering, 2007. 133(2): p. 225-234.
54. Beeby, A., R. Scott, and A. Jones, Revised code provisions for long-term deflection calculations.
Proceedings of the ICE-Structures and Buildings, 2005. 158(1): p. 71-75.
55. Malvar, L., J. Cox, and K.B. Cochran, Bond between carbon fiber reinforced polymer bars and
concrete. I: Experimental study. Journal of composites for construction, 2003. 7(2): p. 154-163.
56. Mitchell, D.W. and H. Marzouk, Bond characteristics of high-strength lightweight concrete.
ACI Structural Journal, 2007. 104(1): p. 22.
57. Alavi-Fard, M. and H. Marzouk, Bond of high-strength concrete under monotonic pull-out
loading. Magazine of Concrete Research, 2004. 56(9): p. 545-557.
58. Gupta, A.K. and S.R. Maestrini, Tension-stiffness model for reinforced concrete bars. Journal of
Structural Engineering, 1990. 116(3): p. 769-790.
59. Kankam, C.K., Relationship of bond stress, steel stress, and slip in reinforced concrete. Journal
of Structural Engineering, 1997. 123(1): p. 79-85.
60. Mirza, S.M. and J. Houde. Study of bond stress-slip relationships in reinforced concrete. in ACI
Journal Proceedings. 1979. ACI.
61. Rehm, G., On the fundamentals of steel-concrete bond. Deutscher Ausschuss für Stahlbeton,
1961. 138: p. 1-59.
62. Nilson, A.H. Nonlinear analysis of reinforced concrete by the finite element method. in ACI
Journal Proceedings. 1968. ACI.
63. Martin, H., On the interrelation among surface roughness, bond and bar stiffness in the
reinforcement subject to short-term loading. Deutscher Ausschuss Stahlbeton, 1973. 228: p. 1-
50.
64. Eligehausen, R., E.P. Popov, and V.V. Bertero, Local bond stress-slip relationships of deformed
bars under generalized excitations. 1982.

158
Chapter 4 – Time-dependent behaviour of concrete

65. Huang, Z., B. Engström, and J. Magnusson, Experimental investigation of the bond and
anchorage behaviour of deformed bars in high strength concrete. Goteborg: Division of
Concrete Structures, Chalmers University of Technology, 1996.
66. Harajli, M., M. Hout, and W. Jalkh, Local bond stress-slip behavior of reinforcing bars
embedded in plain and fiber concrete. ACI Materials Journal, 1995. 92(4).
67. Murcia-Delso, J., A. Stavridis, and B. Shing, MODELING THE BOND-SLIP BEHAVIOR OF
CONFINED LARGE-DIAMETER REINFORCING BARS. COMPDYN 2011 III ECCOMAS
Thematic Conference on Computational Methods in Structural Dynamics and Earthquake
Engineering M. Papadrakakis, M. Fragiadakis, V. Plevris (eds.) Corfu, Greece, 25–28 May,
2011.
68. Barbosa, M., Evaluation of the behavior of the bond in ordinary and high strength concrete.
2001, Doctoral Thesis, COPPE/UFRJ (in Portuguese).
69. Oh, B.H. and S.H. Kim, Realistic models for local bond stress-slip of reinforced concrete under
repeated loading. Journal of Structural Engineering, 2007. 133(2): p. 216-224.
70. de Almeida Filho, F.M., K. Mounir, and A.L.H. El Debs, Bond-slip behavior of self-compacting
concrete and vibrated concrete using pull-out and beam tests. Materials and Structures, 2008.
41(6): p. 1073-1089.

159
Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

CHAPTER 5

MECHANISM OF FLEXURAL DEFLECTION


IN REINFORCED CONCRETE STRUCTURES
Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

Table of Contents

5.1. INTRODUCTION .................................................................................................................. 160


5.2. DEFLECTION CONTROL IN SERVICEABILITY DESIGN .............................................. 163
5.3. RATIO OF LONG-TERM TO SHORT-TERM DEFLECTION ........................................... 164
5.4. REAL-SCALE TIME-DEPENDENT DEFLECTION MONITORING IN LITERATURE .. 165
5.5. EXPERIMENTAL MONITORING OF DEFLECTIONS IN LABORATORY CONDITIONS
FROM LITERATURE ............................................................................................................................. 167
5.5.1. Experiment by Uy and Bradford (1994, 1997) ................................................................... 167
5.5.2. Experiment by Gilbert and Guo (2005) .............................................................................. 169
5.5.3. Experiment by Park et al. (2012) ........................................................................................ 171
5.6. MINIMUM THICKNESS PROVISIONS .............................................................................. 173
5.7. EFFECTIVE PARAMETERS IN INSTANTANEOUS DEFLECTION ............................... 175
5.8. EFFECTIVE PARAMETERS IN LONG-TERM DEFLECTION ......................................... 176
5.9. TENSION STIFFENING ....................................................................................................... 177
5.9.1. Existing Models of Tensile Stress Block of Concrete in Literature ................................... 178
5.10. EFFECTIVE MOMENT OF INERTIA.................................................................................. 187
5.11. EXISTING MODELS OF EFFECTIVE MOMENT OF INERTIA ....................................... 193
5.12. SUMMARY............................................................................................................................ 195
REFEERENCES ...................................................................................................................................... 197
Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

List of Figures

Figure 5.1: Long-term deflection of actual structure after stripping [16]................................................. 165
Figure 5.2: Profiled composite slab and beam [27] .................................................................................. 167
Figure 5.3: Test set-up for long-term flexural deflection of beams and slabs [27] .................................. 168
Figure 5.4: Long-term deflections of beams at mid-span [27] ................................................................. 169
Figure 5.5: Loading history of slab specimens [30] ................................................................................. 170
Figure 5.6: Ratio of long-term to instantaneous deflection in concrete slabs [30] ................................... 171
Figure 5.7: Specimens dimension and test set-up for long-term flexural deflection [32] ........................ 172
Figure 5.8: Loading history of the specimens [32] ................................................................................... 172
Figure 5.9: Stress-strain behaviour of concrete in tension and compression zones .................................. 177
Figure 5.10: Equivalent stress-strain relationship for tensioned concrete [49] ........................................ 178
Figure 5.11: Influence of α1 and α2 in the moment-curvature (a, b) and moment-strain (c) laws [50] ..... 179
Figure 5.12: Deformation response of idealized reinforcement at critical section [70] ........................... 187
Figure 5.13: Flexural response of reinforced concrete members ............................................................. 188
Figure 5.14: Average moment versus instantaneous curvature relationship [74] ..................................... 190
Figure 5.15: Theoretical moment-curvature response at the critical section using Bischoff’s method (Eq.
5.1) with different tension stiffening values [76] ............................................................................ 191
Figure 5.16: Deflection comparison using Branson’s (Eq. 5.1) and Bischoff’s expression for Ie [76] ..... 191
Figure 5.17: Effect of tension stiffening on member stiffness [75] .......................................................... 193
Figure 5.18: Effect of cracking moment and tension stiffening on the member response after cracking [1]
......................................................................................................................................................... 193
Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

List of Tables

Table 5.1: Limit state definition and major classifications in codes of practice ...................................... 161
Table 5.2: Real-scaled monitoring of time-dependent deflection of reinforced concrete structures ....... 166
Table 5.3: Minimum thickness provisions for slabs to control the time-dependent deflection ................ 174
Table 5.4: Existing models for tensile stress-strain diagrams of reinforced concrete considering tension
stiffening .......................................................................................................................................... 180
Table 5.5: Existing models of effective moment of inertia in RC section ................................................ 194
Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

Chapter 5: Mechanism of flexural deflection in reinforced concrete


structures

5.1. INTRODUCTION

Deflection is a common concern of almost all design codes and technical drafts in regard to
the time-dependent behaviour of concrete. Short-term and long-term deflections are considered
in serviceability limit states. Due to the nonlinear nature and complicated behaviour of concrete
in the presence of reinforcement, deflection prediction is full of unknown parameters. For
accurate prediction of the deflection, comprehensive information on the time-dependent and
nonlinear material behaviour is required.

There are considerable differences between the predicted values of deflection and the
experimental values in the Reinforced Concrete (RC) members subjected to service loads,
particularly in long-term deflection of concrete structures. Existing codes of practice give an
overall estimation of the long-term deflection as a multiple of the short-term or elastic deflection
of concrete. Such estimation sometimes may lead to an entirely wrong prediction of deflection
and consequently, unsafe or non-economical concrete construction.

The majority of the equations in codes of practice are applied to predict the time-dependent
behaviour of the conventional concrete. However, there is also some relevance to other types of
concrete such as Light-Weight Concrete (LWC). The modern design codes have approved the
limit states method in structural design, in which different limit states or the design
requirements such as adequate strength and serviceability which need to be fulfilled,
simultaneously. The required minimum performance limits are specified for each limit state and
each one may become critical of the design of a particular member [1].

Table (5.1) explains the descriptions and major classifications of the limit state design in
various codes of practice.

160
Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

Table 5.1: Limit state definition and major classifications in codes of practice
Code Limit states description
To predict the strength and serviceability of reinforced and prestressed concrete structures, the structural
ACI 209.2R-

engineer requires an appropriate description of the mechanical properties of the materials, including the
08 [2]

prediction of the time-dependent strains of the hardened concrete. The prediction of shrinkage and creep is
important to assess the risk of concrete cracking, and deflections due to stripping-reshoring.
Serviceability limit state: The design should be such that the structure will not suffer local damage which
would shorten its intended life or incur excessive maintenance costs. In particular, calculated crack widths
should not exceed those permitted in 4.1.1.1
- Cracking. Cracking of concrete should not adversely affect the appearance or durability of the structure.
- Stress limitations. To prevent unacceptable deformations occurring, compressive stresses in concrete and
stresses in steel should be calculated by linear elastic analysis
- Deflections. The deflection of the structure, or any part of it, should not be such as to affect adversely the
BS-5400-4 (1990) [3]

appearance or efficiency of the structure.


Ultimate limit states: The strength of the structure should be sufficient to withstand the design loads, taking
due account of the possibility of overturning or buckling. The assessment should ensure that collapse will not
occur as a result of rupture of one or more critical sections, by overturning or by buckling caused by elastic
or plastic instability, having due regard to the effects of sway when appropriate.
- Rupture or instability. The assessment of the structure under the design loads appropriate to this
limit state should ensure that prior collapse does not occur as a result of rupture of one or more critical
sections, buckling caused by elastic or plastic instability, or overturning.
The effects of creep and shrinkage of concrete, temperature difference and differential settlement need not be
considered at the ultimate limit state provided that these effects have been included in the appropriate load
combinations to check the stress limitations given in 4.1.1.3 for the serviceability limit state.
Ultimate Limit State
Attainment of the bearing capacity of a structural part or the structure as a whole is classified as ultimate
limit state.
The ultimate limit states considered are distinguished by the consequences of their exceedance.
The loss of static equilibrium (Section 1.6.5) when a structural part or the whole structure (considered as
CEB-FIP-90 [4]

rigid bodies) is overturned, is lifted or slides.


Exceeding the resistance of one or more critical regions of the structure.
Serviceability limit states:
The serviceability limit states associated with the general requirements refer to:
limited local structural damage such as excessive cracking or excessive compressive stresses, producing
irreversible strains and micro cracks;
Deformations which produce unacceptable damage in non-structural elements
Vibrations resulting in discomfort, alarm or loss of Utility
Design Limit states: Conditions of a structure at which it ceases to fulfil the designed functions.
1- Durability: Concrete structures shall satisfy the durability requirements of CSA A23.1, CSA A23.4, or
CSA-A23.3-04 [5]

CAN/CSA-S413, as applicable, for the intended use and exposure conditions.


2- Fire resistance: Concrete structures shall satisfy the fire resistance requirements of the applicable
building code.
3- Ultimate limit states: Structures, structural members, and connections shall be designed such that
factored resistance is greater than or equal to the effect of factored loads, with the effect of factored loads

161
Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

Table 5.1: Limit state definitions and major classification in codes of practice (continued)
Code Limit states description
with the effect of factored loads being determined as specified in Clauses 8.2 and 8.3 and the factored
resistance being determined as specified in Clause 8.4.
4- Serviceability limit states
4.1 Deflections: Structures and structural members shall be designed to satisfy the deflection control
requirements of Clauses 9.8 and 13.2.7, with loadings as specified in Clause 8.3.3.
CSA-A23.3-04 [5]

4.2 Local damage and cracking: Structural members and connections shall be designed to meet the
minimum reinforcement area and maximum reinforcement spacing requirements of this Standard as well as
the requirements of Clauses 10.6 and 18.1 to 18.4, with loadings as specified in Clause 8.3.3.
Note: This Standard does not specifically limit crack widths.
4.3 Vibrations: In the design of structures and structural members, consideration shall be given to
controlling vibrations within acceptable limits for the intended use.
5- Structural integrity: Consideration shall be given to the robustness of the overall structural system to
minimize the likelihood of a progressive type of collapse.
SERVICEABILITY LIMIT STATES (SLS)
- stress limitation (section 7.2): The compressive stress in the concrete shall be limited in order to avoid
EN 1992-1-1:2004 (E) [6]

longitudinal cracks, micro-cracks or high levels of creep, where they could result in unacceptable effects on
the function of the structure.
- Crack control (section 7.3): Cracking shall be limited to an extent that will not impair the proper
functioning or durability of the structure or cause its appearance to be unacceptable.
- Deflection control: Deformations should not exceed those that can be accommodated by other connected
elements such as partitions, glazing, cladding, services or finishes. In some cases limitation may be required
to ensure the proper functioning of machinery or apparatus supported by the structure, or to avoid ponding on
flat roofs.
Limit state design:
JSCE (2007)

Safety limit state: Cross sectional failure, fatigue failure, displacement, deformation, mechanisms
[7]

Serviceability limit state: Appearances, vibration, comfort of vehicle ride, etc.


Water tightening, resistance against damage (maintenance of functionality)
The acceptable limit for the safety and serviceability requirements before failure occurs is called a ‘limit
IS 456 : (2000) [8]

state’.
Limit State of Collapse: rupture, buckling due to elastic or plastic instability (including the effects of sway
where appropriate) or overturning. The resistance to bending, shear, torsion and axial loads at every section
Limit States of Serviceability: Deflection, Cracking
Other Limit States

Limit state design:


AS-3600-09

Ultimate limit state


[9]

Serviceability limit state: Deflection, cracking, Vibration

Limit state design :


1170.0(2002)
AS/NZS

Serviceability limit states: deflection , cracking


[10]

Ultimate limit states: For Australia as AS-3600 and For New Zealand: as described in Table 3-3

162
Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

5.2. DEFLECTION CONTROL IN SERVICEABILITY DESIGN

According to limit state definitions in codes of practice and literature, serviceability limit
states can be classified into three categories as described in the following terms:

1. Motion perception (vibration and acceleration): common samples include human


discomfort caused by machinery or the wind, particularly if resonance occurs. Floor vibrations
from people or machinery and acceleration, particularly in tall buildings under wind load are
usual areas of concern in this category;

2. Deterioration: Included are such items as corrosion, weathering, efflorescence,


discoloration, rotting, and fatigue;

3. Deformation (deflection, curvature and cracking): Common examples include local


damage to non-structural elements (e.g., ceilings, cladding, and partitions) due to deflections
under dead, live, wind, or seismic loads, and damage from temperature change, moisture,
shrinkage, or creep effects.

The effects of the excessive deflection on performance of the reinforced concrete structures
range from a compromised performance of the structure to unsightly crack openings within the
roofs and walls. Excessive deflection may cause damage to adjacent parts of the building
(cracking and spalling of brittle partitions) and non-load bearing walls that affect a building’s
appearance (visible sag of floors and ceilings, visible leaning of walls and columns). Also
building usage and function may be influenced by the curvature of floors, and inclinations may
cause problems such as tilting of furniture, movement of trolleys and slipping. [11].

Time-dependent deformations of concrete members depend on many factors including the


properties of concrete, the geometry of the member and ambient conditions such as temperature
and relative humidity. The deflections due to these time-dependent effects can be significant,
especially for cracked beams. Cracking has been commonly observed in RC beams and slabs in
service. In cases where deflection control is critical, it is important to predict the long-term and
time-dependent deflections accurately.

As a general conclusion, deflection of RC flexural member increases with time, mainly due
to:

a) Differential shrinkage (causing differential strains across the cross-section, resulting in


curvature);

b) Creep under sustained loading.

163
Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

Other effective parameters of the deflection of RC structures have been described


comprehensively in Chapter Four of this study.

5.3. RATIO OF LONG-TERM TO SHORT-TERM DEFLECTION

There is no guidance as to when the short-term deflection ends and long-term deflection
begins. The methods of predicting or calculating the deflections are deterministic in nature.
Even, with the most sophisticated methods of analysis using experimentally determined material
properties, the range of variation between the measured and computed results for the
instantaneous (short-term) as well as long-term deflection are high. Branson (1977) [12] showed
that the coefficient of variation for such deflections is of the order of 15% to 20% and higher.
Jokinen and Scanlon (1987) [13] analysed experimental data from several identical slabs in a
multi-story building. They reported a coefficient of variation between the calculated and
measured deflections in the range of 30% for both short-term and long-term deflections.

Generally, to determine the long-term deflection of a concrete member, the time-delayed


effects of shrinkage, creep and the resulting reduction in tension stiffening need to be
considered accurately. Additional factors which can contribute to the increased long-term
deflection include the development of new cracks, widening of the initial cracks, and effects of
the repeated load cycles [14].

Shrinkage induced strain that develops with time causes additional curvature known as
shrinkage warping. Deflections also increase due to the effects of creep in the concrete member.
The tension stiffening effect is responsible for a significant proportion of the concrete member
stiffness. It decreases over time due to the effects of creep and shrinkage strains [15]. Therefore,
the RC member that is essentially uncracked at the stripping time will gradually become more
cracked; the effective moment of inertia will decrease, and this will contribute to the long-term
deflection.

Both compressive and tensile concrete creep under an effective constant sustained load also
contribute to the rate of decreasing the long-term deflection with increasing time for several
years after casting. Taylor and Heiman (1977) [16] showed that the deflection of slabs and
beams increases with age for five to nine years after stripping. The decreasing rate of time-
dependent deflection in their experiment is given in Figure (5.1).

164
Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

30

25

20

Deflection (mm)
15

10

0
0 1 2 3 4 5 6 7 8 9 10
Time from stripping (Year)

Figure 5.1: Long-term deflection of actual structure after stripping [16]

The coefficient of creep and shrinkage (kcs) in AS-3600 (2009) acts as the multiplier of the
instantaneous deflection to calculate the total deflection after a long time. However, using a
single multiplier to predict the time-dependent deformations of concrete members cannot cover
all uncertainties. Gilbert (2001) also noted that the method of multiplying the ‘load induced
short-term deflection’ by a multiplier, especially in slabs is fundamentally wrong [17] because
the shrinkage warping occurs irrespective of the load applied and even occurs in unloaded
concrete slabs and contributes significantly to long-term deflections [15].
The uncertainties may arise from changing the material properties, the geometry of the
member, and the ambient conditions such as temperature and relative humidity cannot yield
consistent and accurate values for the calculated and long-term deflection ([17],[18],[19], [20]).
In this regard, Gilbert (2007) [20] criticised the use of creep and shrinkage factor in AS-3600 as
a ‘ball-park estimate’, that does not take into account the shrinkage and creep properties of the
concrete, the load history, the age at first loading and environmental effects, hence often failing
to adequately predict the long-term deflection.

5.4. REAL-SCALE TIME-DEPENDENT DEFLECTION MONITORING IN


LITERATURE

There are limited numbers of reported in-situ monitoring of deflection behaviour of RC


structures under service loads. Table (5.2) illustrates some case studies in which the deflection
has been measured in the actual conditions, real-scale elements in the real conditions of
temperature, environment effects, changing the loading level and every possible situation during
and after the construction. The type of structure and monitored element, loading duration and a
brief comparison with predictions of codes of practice and laboratory investigations are given in
each case.

165
Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

Table 5.2: Real-scaled monitoring of time-dependent deflection of reinforced concrete structures

Case study Specimen type Duration Conclusions


ACI Code method underestimates
1-Two story commercial building- the long-term deflections of flexural
members.
suspended floor of beams and slabs, -13 days
In shallow members with little or no
flat plate roof compressive reinforcement in the
mid-span region, difference between
real and predicted values increases
2- Three story car park- flat slab -56 days
Heiman and separate calculations of long-
(1974) [21] term deflection due to creep and
3-Fifty story circular building- radial - 25to 32 shrinkage give more accurate
tapered beam floor days results.
Existing codes don’t consider the
heavy construction loads that
4-Four story motel building- flat slab -28days strongly affect the long-term
floor deflection.
Lab tests-based calculations may not
Sydney, Australia
be sufficient to predict the real
behaviour of RC elements.
ACI-318 does not adequately
account for the early-age high
Two low-rise (3 story) reinforced
construction loads and the time-
Kaminetzky and concrete buildings 10 dependent cracking.
Stivaros (1992) [22] slabs have been overloaded during months Most of the estimated deflections for
construction, USA the various slab areas do not
compare well with the measured
values.
Vollum et al. (2002) Reliance on span/ depth ratio may
[23], be insufficient.
St George Wharf case study Longer spans and more slender flat
European Concrete 5 months
- slab deflection, Cardington, UK slab structures are more important
Building projects, regarding the serviceability limit
(2004) [24] states of deflection and vibration.
Precast Composite Slab Span System
(PCSSS), implemented for short to
French et al. (2010) 24 Cracking and strain distribution are
moderate span bridges (6-15 m.
[25] months more uniform in lab conditions.
range).
Minnesota-USA
Considering a cracked slab as
uncracked is a significant danger. If
the factors that contribute to
cracking (load history,
1- West Washington Street building
column/middle strip effects,
(No.215) restrained shrinkage, shrinkage
Hirsch et al. (2012) 2-Fifty story building- 2way concrete warping) cannot be included directly
35 days into the calculation, conservative
[26] flat slab
adjustments need to be made to the
3-Aqua high-rise project, 82 story calculations to ensure that cracking
Chicago , USA that occurs is detected.
Caution should be exercised in span
to depth ratios for preliminary
design of members with unusually
high loads or irregular load histories.

166
Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

5.5. EXPERIMENTAL MONITORING OF DEFLECTIONS IN


LABORATORY CONDITIONS FROM LITERATURE

In the real-scaled case studies of the real structures presented in Table (5.2), the recoding
period of deflection and time-dependent parameters are too short and could not be extended to
the whole service life of the RC structure. They monitor just a specified duration of the service
life, while in the RC structures, due to the nonlinear and complex behaviour of composite
materials, the curing, construction and loading conditions before the monitoring time strongly
affect the time-dependent behaviour. These effects may sometimes significantly change the
predictions based on the laboratory investigations and the relationships developed in design
codes.

5.5.1. Experiment by Uy and Bradford (1994, 1997)


Uy and Bradford (1994) [27] and Uy (1997) [28] conducted experiments to investigate the
time-dependent behaviour of simply supported profiled composite beams for 450 days. The
results were compared with the predictions of BS-5950 [29] to verification purposes.

The profiled composite slab in a steel framed building and a profiled composite beam and
slab in a concrete framed building are shown in Figure (5.2). The flexural test set-up is
presented in Figure (5.3).

Profiled composite slab in a steel-framed building Profiled composite beam and slab in a concrete framed building
Figure 5.2: Profiled composite slab and beam [27]

167
Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

Figure 5.3: Test set-up for long-term flexural deflection of beams and slabs [27]

Figure (5.4) shows the long-term deflection of beams over 450 days. The beams Rl and P1
were subjected to self-weight only to study the effect of shrinkage deformations on the
behaviour. The beams R2 and P2 were subjected to self-weight with an additional 5 KN/m
sustained load to study the effect of shrinkage and creep deformations.

According to Figure (5.4), the long-term deflection after 450 days is about 3 to 4 times
greater than the instantaneous deflection in all the specimens. The increasing trend of deflection
in the beams subjected to self-weight and additional sustained load are slightly increasing over
time especially after 100 days of first loading. In other words, the distance between the pair of
curves is increasing by time. It is due to the increasing effect of creep strain under the higher
sustained load in the beams subjected to higher loading level.

Other conclusions from their study are described as following:

- The span to depth ratio is not adequate to limit the long-term deflection of composite
beams and slabs;
- The beneficial effect of lateral restraints of supporting beams on the composite slab
stiffness and long-term behaviour are ignored in design of these types of structures.

168
Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

Figure 5.4: Long-term deflections of beams at mid-span [27]

5.5.2. Experiment by Gilbert and Guo (2005)


Gilbert and Guo (2005) [30] performed a three-year experimental program by testing of 7
specimens of large-scale reinforced concrete flat slabs. Overall plan dimensions of the slabs
were 6200 mm by 7200 mm continuous over two spans in each direction supported by columns.
Applied different environmental conditions and loading history to the specimens were as
illustrated in Figure (5.5).

Deflections of slabs were recorded at different ages over the 750 days testing period. The
ratio of deflection in different ages to instantaneous deflection after loading is presented in
Figure (5.6).

Generally, the long-term to short-term (instantaneous ) deflection ratio may be as large as


two to three [14], according to the design codes. However, the ratios in this experimental
program vary in the range of 5 to 9 at the end of the test, or at the age just before the loading
blocks were removed. The ratios shown in Figure (5.6) are recorded under different conditions
of loading histories and wetting and drying periods. These high ratios in various slabs depend
on the loading history and various curing conditions.

169
Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

Figure 5.5: Loading history of slab specimens [30]

170
Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

Figure 5.6: Ratio of long-term to instantaneous deflection in concrete slabs [30]

Gilbert and Guo (2005) also investigated the effect of wetting and drying cycles on the
deflection behaviour of slabs. They found that the immediate effect of wetting after a long
period of drying is comparable to the effect of the full-service load and in-service behaviour of
the slabs which are very sensitive to the degree of exposure (15% reduction of the ratio in slab
S2 after 3 months). The duration and magnitude of warping due to drying shrinkage is directly
proportional to the free drying shrinkage of the concrete. It continues at a decreasing rate with
increasing time for several years after casting and is independent of loading [31]. For both slabs
S1 and S2, within 24 hours after the top surface was thoroughly wetted, the deflection was
reduced significantly due to the redistribution of internal actions.

According to Figure (5.6), deflection in all slabs was rapidly extended in first ages and
continued with slow growth without stopping even after 750 days. This conclusion confirms the
test results obtained by Taylor and Heiman (1977) [16] previously.

5.5.3. Experiment by Park et al. (2012)


Park et al. (2012) [32] investigate the instantaneous and long-term deflections of RC slabs
subjected to construction load at their early ages. The concrete was cured at low-temperature
and the loading was sustained for 291 days after initial loading. The test program included
seven one-way slabs and one two-way slab.

171
Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

Dimensions of the simply supported one-way slab were 4500 mm length ¯800 mm width
¯160 mm depth. Two concentrated loads were applied at the mid-span. The shape of two-way
slab was cruciform, which was vertically supported at four ends. Figure (5.7) shows the section
and dimensions of the specimens and the testing set-up view.

Figure 5.7: Specimens dimension and test set-up for long-term flexural deflection [32]

They studied the effect of changing load value during the test on the deflection changes in
the specimens. Figure (5.8) shows the loading history of the specimens for one-step and multi-
step loading cases.

Single step loading (uniform sustained loading)

Single step loading (uniform sustained loading)

Figure 5.8: Loading history of the specimens [32]

172
Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

The following items were concluded from their study:

- The lower curing temperature strongly reduces the compressive strength and modulus
of elasticity of concrete by 30%;
- The early age loading at three days increases the instantaneous deflection by 31%;
- The ratio of long-term deflection and total deflection to the instantaneous deflection
after 110 days were in the range of 1 to 2 and 2 to 3 in different slabs respectively. In
other words, the time-dependent deflection multiplier was affected by loading history,
curing condition and initial loading age. These ratios also are being affected by the ratio
of the applied service moment to the cracking moment (Ma /Mcr).

5.6. MINIMUM THICKNESS PROVISIONS

There are different simplified methods in design codes to control the deflection and to ensure
the serviceability of flexural RC elements. Limiting the minimum thickness is one of these
simple methods of deflection control. This limit is just a general estimation of the deflection and
is not able to guarantee the expected behaviour under different conditions of loading, material
technology and environmental effects. Table (5.3) summarize the limitations proposed in
various codes of practice to control the deflection of RC elements. The limitation suggested by
Lee and Scanlon (2010) [33] is presented in the Table (5.3) also.

173
Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

Table 5.3: Minimum thickness provisions for slabs to control the time-dependent deflection

Code Minimum thickness of non-prestressed one-way slabs unless deflections are calculated
Simply supported One end continuous Both ends continuous Cantilever
l/20 l/24 l/28 l/10
Minimum thickness of two-way slabs without interior beams
Without drop panels
Exterior panels Interior panels
fy,psi Without edge beams With edge beams -
60,000 ln /30 ln /33 ln /33
ACI 318-08 [34]

l is span length of one-way slab , ln is length of clear span measured face-to-face of supports
For slabs with beams spanning between supports on all sides
αfm ≤ 0.2 Slabs without drop panels - 5 in. and Slabs with drop panels - 4 in.
݂௬
݈௡ ሺͲǤͺ ൅ ሻ
0.2 ≤ αfm ≤ 2 ݄ൌ ʹͲͲͲͲͲ ‘–Ž‡••–Šƒͷ‹Ǥ
͵͸ ൅ ͷߚሺߙ݂௠ െ ͲǤʹሻ
݂௬
݈௡ ሺͲǤͺ ൅ ሻ
αfm > 2 ݄ൌ ʹͲͲͲͲͲ ‘–Ž‡••–Šƒ͵Ǥͷ‹Ǥ
͵͸ ൅ ͻߚ
αfm is average value of αf for all beams on edges of panel
αf is ratio of flexural stiffness of beam section to flexural stiffness of width of slab
Lef is effective span, taken as less of (ln + d) and l;
d is effective depth of cross section;
Δ /Lef is deflection limit;
Fd.ef is effective design load, per unit area;
‫ܮ‬௘௙
k3 = 1.0 for one-way slab, rectangular slabs supported on four sides
AS 3600-09 [9]

݀
ଵȀଷ = 0.95 for two-way flat slab without drop panels;
൬οൗ‫ ܮ‬൰ ‫ܧ‬௖ k4 is deflection constant which may be taken as:
௘௙
ൌ ݇ଷ ݇ସ ൦ ൪
‫ܨ‬ௗǤ௘௙ (a) for simply supported slabs, 1.6; or
(b) for continuous slabs, where in adjoining spans ratio of longer span to shorter
span does not exceed 1.2 where no end span is longer than an interior span-(i) 2.0
in an end span; or (ii) 2.4 in interior spans; and (c) for edge-supported slabs k4
varies from 1.7 to 2.0 (Table 9.3.4.2 in AS 3600-2001).
ሺͶ͹͹ െ ݂௦ ሻ
Basic span/effective depth ratios for rectangular Modification factor for ൌ ͲǤͷͷ ൅
‫ܯ‬
ͳʹͲሺͲǤͻ ൅ ௨ଶ ሻ
section tension reinforcement : ܾ݀
൑ ʹǤͲǡ ݂௦ ൌ ʹȀ͵݂௬
Support conditions Rectangular section Mu is design ultimate moment at centre of span or,
for cantilever, at support;
BS 8110-1-1997 [35]

fs is estimated design service stress in tension


reinforcement;
Cantilever 7 fy is yield strength of reinforcement;
b is effective width of rectangular beam;
d is effective depth;
Simply supported 20
For spans exceeding 10 m, Table 3.9 should be
multiplied by 10/span, except for cantilevers
Continuous 26 where the design should be justified by
calculation.
For flat plate, span/effective depth ratio should be
multiplied by 0.9.

174
Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

Table 5.3: Minimum thickness provisions for slabs to control the time-dependent deflection (continued)

Code Minimum thickness of non-prestressed one-way slabs unless deflections are calculated
݈ ߩ଴ K is factor to take into account different structural
ൌ ‫ ܭ‬൤ͳͳ ൅ ͳǤͷඥ݂௖௞
݀ ߩ systems: (a) simply supported 1.0; (b) one end
If ρ ≤ ߩ଴
൅ ͵ʹඥ݂௖௞ ሺ continuous 1.3; and (c) both end continuous 1.5;
ρο ߩ
ଵ଴షయ
ଷȀଶ ρo is reference reinforcement ratio, ൌ ඥ݂௖௞ ;
െ ͳሻ ൨
Eurocode 2-04 [6]

ρ is required tension reinforcement ratio at mid-span to


resist moment due to design loads (at support for
݈ ߩ଴ cantilevers);
ൌ ‫ ܭ‬቎ͳͳ ൅ ͳǤͷඥ݂௖௞
݀ ߩ െ ߩᇱ ρ′ is required compression reinforcement ratio at mid-
If ρ >
span to resist moment due to design loads (at support
ρο
ͳ ߩᇱ
൅ ඥ݂௖௞ ඨ ቏ for cantilevers);
ͳʹ ߩ଴
fck is specified compressive strength of concrete, in
MPa; and l is span length.
ଵȀଷ
݈௡ ο௜௡௖ ʹͶͲͲ‫ܭ‬஽௉ ‫ܧ‬௖ ሺܾൗͳʹሻ Ws is sustained load (psf [slabs]; plf [beams]); (Pa
ൌ ߚ ൥൬ ൰ ൩
݄ ݈ ௔௟௟௢௪ Ɉ݇஺ோ ݇௦௦ ሺɉܹ௦ ൅ ܹ௅ ሺܽ݀݀ሻሻ [slabs]; N/m [beams]);
(U.S. Customary Units) WL(add) is additional live load (psf [slabs]; plf
[beams]); (Pa[slabs]; N/m [beams]);
β = 1, except β is long span/short span ≤ 2.0 for edge-
supported slabs;
κ is deflection coefficient depending on support
condition: equals
Lee and Scanlon (2010) [33]

5 for simply supported, 1.4 for both ends continuous,


2 for one end continuous, and 48 for fixed end
cantilever;
ଵȀଷ
݈௡ ο௜௡௖ ͲǤͲͳ͸͹‫ܭ‬஽௉ ‫ܧ‬௖ ܾ
ൌ ߚ ൤൬ ൰ ൨ kDP = 1, except kDP = 1.35 for slab with drop panels;
݄ ݈ ௔௟௟௢௪ Ɉ݇஺ோ ݇௦௦ ሺɉܹ௦ ൅ ܹ௅ ሺܽ݀݀ሻሻ
kSS = 1, except kSS = 1.35 for column supported two-
way slab systems;
(SI Units)
kAR = 1, except kAR = 0.2 + 0.4β for edge-supported
slabs;
b = 12 in. (1000 mm for SI) for one- and two-way
slabs = beam width (= web width, bw for T-beams) (in.
for U.S. Customary Units and mm for SI Units);
(Δinc)allow is required incremental deflection limit; and
λ is long-time multiplier for sustained loads (ACI 318,
Section 9.5.2.5).

5.7. EFFECTIVE PARAMETERS IN INSTANTANEOUS DEFLECTION

The principle factors which affect the initial or short-term deflection of RC flexural members
under service loads are the modulus of elasticity, load distribution and support conditions. Other

175
Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

parameters are variable cross-section and hence, variable moment of inertia, load level and the
degree of cracking along the member.

Cracking and tension-stiffening parameters probably have the most significant effect on
numerical results of concrete members subjected to short-term loading. The free shrinkage
strain of concrete may be of a magnitude well-exceeding the cracking strain, even at first
loading. The shrinkage might significantly affect the cracking resistance and deformations of
RC elements subjected to short-term loading. Bischoff (1983), Gilbert (2001), Kaklauskas and
Gribniak (2005), Gribniak et al. (2008), Gribniak (2009) and Kaklauskas et al. (2009) discussed
the necessity to assess the shrinkage influence on short-term deformations of RC members [36].

Obviously, for a cracked member, assuming every cross-section is uncracked leads to


underestimated deformation and assuming that is going to be fully cracked leads to even
sometimes grossly overestimated deformation [37].

5.8. EFFECTIVE PARAMETERS IN LONG-TERM DEFLECTION

The time-dependent nature of the effective parameters and the interaction between the
parameters, along with the unknown properties of concrete at the design stage and the lack of
experimental data to compare the results of the analysis with different techniques and codes,
make the deflection prediction of RC members, particularly slabs a challenging problem.
Mechanical properties of concrete affect the deflection directly by affecting the structural
stiffness of the element and indirectly to define the moment redistribution due to cracking. So
implementing the accurate mechanical properties of concrete and incorporating uncertainties is
a crucial factor in deflection calculations.

Several models for the non-linear and time-dependent analysis of RC flexural members have
been developed, which take into account the non-linearity of materials and the effects of creep
and shrinkage of concrete. Sectional analyses with different levels of accuracy have been
performed by Clarke et al. (1988) [38], Gilbert (1988) [39], Murcia (1991) [40], Favre and
Charif (1994) [41], Nie and Cai (2000) [42], Chiorino (2005) [43] and Ghali et al (2006) [44].
At the structural level, general methods based on the layer filament beam elements have been
developed by Kang and Scordelis (1980) [45], Marí (1984) [46], Elbadry and Ghali (1985) [47],
Cruz (1998), and Ulm (1994) [48]. These methods include the non-linearity of materials such
as cracking, yielding, crushing and load reversals, second order effects, time-dependent effects
of creep and shrinkage and their interactive effects.

176
Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

5.9. TENSION STIFFENING

Tension stiffening is attributed to the fact that concrete does not crack suddenly and entirely
but undergoes progressive micro-cracking (strain softening). After first cracking in the RC
member under tension or flexure, the intact concrete between adjacent primary cracks takes the
significant portion of the tensile force, mainly in direction of the reinforcement. This is the
result of the existing bond between the steel and the surrounding concrete.

Generally in the design and evaluation of RC elements, concrete poses no tension capacity
after cracking in the tensile zone. Owing to the effect of tension stiffening, the bending stiffness
of the member is considerably higher than that based on a fully cracked section, where concrete
in tension is assumed to take zero stress. The average tensile stress in the concrete is a major
percentage of the tensile strength of concrete. Figure (5.9) shows comparative diagrams of the
compression and tension capacity of concrete regarding the stress-strain relationship. These
empirical models vary in the applied techniques, properties of the materials and loading age as
input data to ascertain the creep strain changes by time and related creep coefficient.

Figure 5.9: Stress-strain behaviour of concrete in tension and compression zones

This tension stiffening effect may be substantial in the performance of beams and slabs
under service load. This phenomenon is also effective in long-term behaviour of concrete
members due to the creep and shrinkage of the concrete.

Various methods have been proposed to implement the tension stiffening effect in the
analysis of RC structures. To demonstrate that an area of concrete located at the level of tensile
steel is effective in providing stiffening is an approach to modelling the tension stiffening.

There are various diagrams to show the tensile behaviour of concrete proposed by
researchers. Damjanic and Owen (1984) [49] proposed an equivalent uniaxial constitutive

177
Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

model for the tensioned concrete model depending only on dimensionless coefficients α1 and α2
as shown in Figure (5.10).

Figure 5.10: Equivalent stress-strain relationship for tensioned concrete [49]

Torres et al. (2004) [50] investigated the effect of these factors on the existing instantaneous
moment-curvature (M-κ) models illustrated in Figures (5.11 a, b) and moment–strain diagram
shown in Figure (5.11c) for a rectangular section. The reference fibre selected for the strain ε0 is
the one corresponding to tensioned steel. They concluded the significant effect of these factors
on the shape of the M–κ and moment– strain laws. In particular, the main effect of α1 is on the
part of the laws that immediately follow the cracking initiation and considerable influence of α2
in the upper part of the diagram (where the applied service moment is considerably greater than
the cracking moment, i.e., Ma>> Mcr). Furthermore, the pairs of α1 and α2 values yielding best
adjustments for the M– κ and the moment–strain laws are different.

5.9.1. Existing Models of Tensile Stress Block of Concrete in Literature


Table (5.4) shows some existing models in the literature to predict the stress-strain
relationship of concrete in tensile zone. The models are mainly proposed based on the
experimental investigations of the sections under direct tension or flexural bending loading.
Some models are developed based on numerical investigations and modification of the existing
models for particular situations.

178
Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

Figure 5.11: Influence of α1 and α2 in the moment-curvature (a, b) and moment-strain (c) laws [50]

179
Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

Table 5.4: Existing models for tensile stress-strain diagrams of reinforced concrete considering tension stiffening

Reference Tensile stress-strain model Notes

Layered analysis of RC section


Scanlon-Murray ݂௧ᇱ : tensile strength of concrete
(1974) [51] εcr: Strain at peak stress ݂௧ᇱ

Ascending branch : linear


Lin and Scordelis
Descending branch : Polynomial
(1975) [52]
Perfect steel-concrete bond assumption

For coefficients refer to Alih and Khelil (2012 ) [54]


Vebo and Ghali
Pt Rt St Ft
model-1977 [53]
Concrete layer adjacent to bar 0.9 0.45 2.2 12.2

fbf and ebf tensile stress and strain at bottom face concrete
Scott ft: concrete ultimate tensile stress
(1983) [55] (EI)u : EI for the uncracked section
(EI)cr : EI for the fully cracked section

180
Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

Table 5.4: Existing models for tensile stress-strain diagrams of reinforced concrete considering tenssion stiffening (contineud)

Reference Tensile stress-strain model Notes

Gilbert and Warner


(1978) [56]
For coefficients refer to Alih and
Khelil (2012 ) [54] in this table
Pt Rt St Ft
Concrete layer adjacent to bar 0.8 0.4 4 10
one layer off the bar 0.5 0.2 4 10
two layers off the bar 0.2 0.1 4 10

ft: Tensile strength of concrete


εct: Strain at peak stress ft
ߝ
‫ۓ‬ ‫ܧ‬௖଴ ߝ݂݅ ൑ͳ
ߝ௖௧
Damjanic and ۖ
ۖߙଵ ݂௧ ሺߙଶ ߝ௖௧ െ ߝሻ
Owen- (1984) [49] ݂௧ ൌ ݂݅ߝ௖௧ ൏ ߝ ൑ ߙଶ ߝ௖௧
‫ ۔‬ሺߙଶ ߝ௖௧ െ ߝ௖௧ ሻ
ۖ
ۖ
‫ە‬ Ͳ݂‫ߙݎ݋‬ଶ ߝ௖௧ ൏ ߝ
Ƚଵ ൌ ͲǤͷǡ ͷ ൑ Ƚଶ ൑ ͳͲ

181
Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

Table 5.4: Existing models for tensile stress-strain diagrams of reinforced concrete considering tension stiffening (contineud)

Reference Tensile stress-strain model Notes

ft: tensile strength of concrete


εct: Strain at peak stress ft
εmax = 20 εct (α1=1 , α2=20 )
ߝ
Schnobrich ‫ۓ‬ ‫ܧ‬௖଴ ߝ݂݅ ൑ͳ
ߝ௖௧
(1985) [57] ۖ
ۖ ߙଵ ݂௧ ሺߙଶ ߝ௖௧ െ ߝሻ
݂௧ ൌ ݂݅ߝ௖௧ ൏ ߝ ൑ ߙଶ ߝ௖௧
‫ ۔‬ሺߙଶ ߝ௖௧ െ ߝ௖௧ ሻ
ۖ
ۖ
‫ە‬ Ͳ݂‫ߙݎ݋‬ଶ ߝ௖௧ ൏ ߝ

ߝ௧ ߝ௧
‫ۓ‬ ݂௧ᇱ ቆ ᇱ ቇ ݂݅ ᇱ ൑ ͳ
ۖ ߝ௧ ߝ௧
ۖ
ߝ
݂௧ ൌ ߚ௧ ݂௧ᇱ ൬ ௧ᇱ ൰
‫۔‬ ߝ௧ ߝ௧
݂݅ ᇱ ൐ ͳ
ۖ
ۖ ߝ ఉ ೟ ߝ௧
Carriera and Chu (1986) ߚ െ ͳ ൅ ൬ ௧ᇱ ൰
and ‫ ە‬௧ ߝ௧
Prakhya and Morely ‫ܧ‬௖௜
ߝ௠௔௫ ൌ ߙߝ௧ᇱ ߙ ൌ ൅ͳ
(1990) [58] ‫ܧ‬௧
ͲǤͶͺ͵‫ܧ‬௖௜
‫ܧ‬௧ ൌ
ͲǤ͵ͻ͵ ൅ ݂௧ᇱ
ߚ௧ : Empirical factor
βt =1 perfect plastic , βt = ∞ sudden loss of strength
ߝ௧ᇱ : strain at peak stress ݂௧ᇱ

RC elements subjected to shear


݂௖௥
Vecchio and ݂௖ଵ ൌ
Collins (1986) [59] ͳ ൅ ඥʹͲͲ߳ଵ
fc1, ε1: principal tensile stress and strain in concrete
fcr: cracking stress

182
Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

Table 5.4: Existing models for tensile stress-strain diagrams of reinforced concrete considering tension stiffening (contineud)

Reference Tensile stress-strain model Notes

‫ܧ‬௖ ൌ ͵͵ʹͲඥ݂௖ᇱ ൅ ͸ͻͲͲሺ‫ ܫܥܣ‬െ ͵͸͵ െ ͺͶሻ


మൗ

݂௧ᇱ ൌ ͲǤ͵ʹͶ ඥ݂௖ᇱ ሺƒ’Šƒ‡ŽሺͳͻͺͶሻ
݂௧ᇱ
Massicote et al. ߝ௖௥ ൌ
‫ܧ‬௖
(1990) [60] ଵ ଵ
‫ܧ‬ଵ ൌ ǡ‫ܧ‬ଶ ൌ ‫ܧ‬௖
଺ ଵ଴
௙೟ᇲమ
ܹ௙ ൌ ͷ cracking energy
ா೎
Total area = 10 ¯ (ascending branch area)

BS 8110-1
(1997) [35]

ߝഥ௧ ͳ ൅ ߚሺͳ െ ܽሻȀܽ


ߪ௧ ൌ ܽǤ ߪ௖௥ ቆͳ െ െ ቇ
ߚ ߚሺߝഥ௧ ሻ௕
ߝ௧ ߪ௖௥
ߝഥ௧ ൌ Ǣߝ௖௥ ൌ
ߝ௖௥ ‫ܧ‬௖
Kaklauskas ܽ ൌ ͲǤ͸ʹͷǡ ܾ ൌ ͳǤͲ
(1999) [61] ߚ ൌ ͵ʹǤͺ െ ʹ͹Ǥ͸ߩ ൅ ͹Ǥͳʹߩଶ Ǣ ߩ ൏ ͲǤͲʹ
ߚ ൌ ͷǢ ߩ ൒ ͲǤͲʹ


ߪ௖௥ ൌ ͲǤʹ͵ට݂௖௨ିଵହ଴ ሺ‫ܽܲܯ‬ሻ
ହǤହൈ௙೎ೠషభఱబ
‫ܧ‬௖ ൌ ൈ ͳͲସ (MPa)
ଶ଻ା௙೎ೠషభఱబ

183
Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

Table 5.4: Existing models for tensile stress-strain diagrams of reinforced concrete considering tension stiffening (cntineud)
Reference Tensile stress-strain model Notes
Applicable for deep beams and walls (plane stress elements)
Based on model of Petersson (1981) [63]
ͳ ʹ ͳͺ ‫ܧ‬௖ ‫ܩ‬௙
ߙଵ ൌ ǡ ߙଶ ൌ ߙଷ ൅ ߙଵ ǡ ߙଷ ൌ
͵ ͻ ͷ ݈௖௛ ݂௖௧ଶ
Foster and Marti and Hillerborg et al. (1976) [64]
(2003) [62] ʹ‫ܧ‬௖ ‫ܩ‬௙
ߙଵ ൌ Ͳǡ ߙଶ ൌ ߙଷ ൌ
݈௖௛ ݂௖௧ଶ
fct : tensile strength of the concrete
Gf: fracture energy;
lch :characteristic length of the finite element

Short-term loading
Modification of ICE, TN-372 model (1983)
Beeby et al.
or model proposed by Scott (1983)
(2005) [65]
ft : tensile strength of the concrete

Long-term loading
Beeby et al.
Modification of ICE, TN-372 model (1983)
(2005) [65]
or model proposed by Scott (1983)

General and fibre reinforced section


Post-peak cyclic loading applied
Nayal and Cyc.1 Cyc.2 Cyc.3 Cyc.4 selected
Rasheed Pt 0.6,0.8,1 0.8 0.8 0.8 0.8
(2006) [66] Rt 0.45 0.3,0.45,0.6 0.45 0.45 0.45
St 4 4 4,6,8 4 4
Ft 10 10 10 5,10,40,100 10

184
Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

Table 5.4: Existing models for tensile stress-strain diagrams of reinforced concrete considering tension stiffening (contineud)

Reference Tensile stress-strain model Notes

fct : concrete tensile strength


εcr : strain at fct

ିఈቀ ቁ
Stramandinoli ߪ௖௧ ൌ ݂௖௧ Ǥ ݁ ఌ೎ೝ
et al. (2008) [67] ߙ ൌ ͲǤͲͳ͹ ൅ ͲǤʹͷͷሺ݊ߩሻ െ ͲǤͳͲ͸ሺ݊ߩሻଶ ൅ ͲǤͲͳ͸ሺ݊ߩሻଷ
‫ܣ‬௦௧ ‫ܧ‬௦
ߩൌ ǡ݊ ൌ 
ܾ݀ ‫ܧ‬௖

General and fibre reinforced section


݂௧ᇱ ൌ ߙଵ ݂௧ Ǣߙଵ ൏ ͳ
݂௧
ߝ௧௨ ൌ ߙଶ ߝ௖௧ ǡߙଶ ൐ ͳǡߝ௖௧ ൌ
‫ܧ‬௖଴
Ng et al. α1 , α2=0.4 and 18 for point load
(2010) [68] α1 , α2=0.5 , 14 for distributed load
ߪ ൌ ‫ܧ‬௖଴ ߝ݂‫ ߝݎ݋‬൑ ߝ௖௧
ߙଵ ݂௧ ሺߙଶ ߝ௖௧ െ ߝሻ
ߪൌ ݂‫ߝݎ݋‬௖௧ ൏ ߝ ൑ ߙଶ ߝ௖௧
ሺߙଶ ߝ௖௧ െ ߝ௖௧ ሻ
ߪ ൌ Ͳ݂‫ߙݎ݋‬ଶ ߝ௖௧ ൏ ߝ
Inoxydable steel reinforcing bar
ܲ௧ ൌ ͲǤͺǡ ܴ௧ ൌ ͲǤͶͷǡ ܵ௧ ൌ Ͷǡ ‫ܨ‬௧ ൌ ͳͲ
଴Ǥ଺଻ ݂௧ᇱ
݂௧ᇱ ൌ ͲǤ͵݂௖௞ ǡߝ௖௥ ൌ
Alih and Khelil ‫ܧ‬௖
(2012 ) [54] ߪ஼ǡ்ௌா ൌ ߩ௘௙௙ ߪௌǡ்ௌா ߪௌǡ்ௌா ൌ ൫ߪௌ െ ߪௌǡூூ ൯
‫ܣ‬௦
ߩ௘௙௙ ൌ
‫ܣ‬௖ǡ௘௙௙
Ac,eff: effective zone of concrete around the rebar

185
Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

Table 5.4: Existing models for tensile stress-strain diagrams of reinforced concrete considering tension stiffening (contineud)

Reference Tensile stress-strain model Notes

Short-term loading
‫ܣ‬௦௧ൗ
ߙ ൌ ͲǤͶͷǡ ߩ ൌ ܾ݀
Bacinskas et al.
( 2012) [69] ݂௖௧௠ ൌ ‫݄ݐ݊݁ݎݐݏ݈݁݅ݏ݊݁ݐ‬From EC-2
ߝ௨௟௧ ൌ ߝ௖௥ ሺ‫ݐ‬଴ ሻ ൈ ሺ͵ʹǤͺ െ ʹ͹Ǥ͸ߩ ൅ ͹Ǥͳʹߩଶ ሻǢ ߩ ൑ ͲǤͲʹ
ߝ௨௟௧ ൌ ͷǢ ߩ ൐ ͲǤͲʹ

Long-term (sustained )loading


‫ܣ‬௦௧ൗ
ߙ ൌ ͲǤͶͷǡ ߩ ൌ ܾ݀
Bacinskas et al.
( 2012) [69] Reduced tensile strength under sustained loading:
݂௖௧ ሺ‫ݐ‬ǡ ‫ݐ‬଴ ሻ ൌ ݂௖௧௠ ሺ‫ݐ‬଴ ሻሾͲǤ͹ͻͶ െ ͲǤͲ͸ ൈ ݈‫݃݋‬ଵ଴ ሺ‫ ݐ‬െ ‫ݐ‬଴ ሻሿ
݂௖௧ ሺ‫ݐ‬ǡ ‫ݐ‬଴ ሻ ൒ ߙ݂௖௧ ሺ‫ݐ‬଴ ሻ

186
Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

5.10. EFFECTIVE MOMENT OF INERTIA

When a reinforced concrete element cracks, its stiffness does not suddenly change to that of
a section where the tension-concrete can be fully disregarded [70]. In fact, the sections where
the cracks are localized are separated by regions where the concrete in tension is uncracked with
an increase of the stiffness. Behaviour at the cracked section used to compute the cracking
moment of inertia (Icr) is assumed to be linear elastic, and nonlinearity of the member stiffness
is taken into account with an effective moment of inertia (Ie) that models the transition from a
gross (uncracked) moment of inertia (Ig) to the cracked transformed moment of inertia (Icr).
Figure (5.12) shows the cracked and uncracked sections of a flexural reinforced concrete
member under sustained loading.

Figure 5.12: Deformation response of idealized reinforcement at critical section [70]

Figures (5.13 a, b) also show the load-deflection and moment-curvature diagrams of a


simply supported RC flexural elements subjected to point-load at the mid-span.

187
Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

a) Load-deflection b) Moment-curvature
Figure 5.13: Flexural response of reinforced concrete members

Various empirical models have been used to model this behaviour, primarily to predict and
check deflections of RC flexural members with code limits. Such models are normally
compatible with moment-curvature (M-κ) space rather than stress-strain (σ-ε) space, so do not
model individual cracks [70].

Branson (1963, 1968) developed Equation (5.1) to predict the curvature and deflections by
the effective moment of inertia directly from elasticity theory. This model was adopted by the
deflection committee of the American Concrete Institute (ACI-435) in 1966 and then in ACI-
318-77 (and recent revisions) building code and Australian code for concrete structures design
(AS-3600-09) with some modifications [71].

‫ܯ‬ ‫ܯ‬
‫ܫ‬௘ ൌ ሺ ௖௥ൗ‫ ܯ‬ሻଶ ‫ܫ‬௚ ൅ ቀͳ െ ሺ ௖௥ൗ‫ ܯ‬ሻଶ ቁ ‫ܫ‬௖௥ ‫ݍܧ‬Ǥ ሺͷǤͳሻ
௔ ௔

Where; Ig, Ie , Icr, Mcr and Ma are representatives for gross moment of inertia, the effective
moment of inertia, moment of inertia of cracked section, cracking moment and service moment
of RC section respectively.

Flexural behaviour after cracking is more accurately reflected by taking a weighted average
of flexibility instead of stiffness, and this leads to a subtle change in Branson’s original
expression. According to an investigation by Bischoff and Gross (2010) [72], in many cases
there is not much difference between using the stiffness at the critical section to represent the
member stiffness or taking into account the change in stiffness along the member length. This is
particularly evident for steel reinforced concrete beams where stiffness of the cracked section
(EcIcr) is typically one-third or more of the uncracked stiffness (EcIg), and the service load
moment (Ma) is less than three times of the cracking moment of section (Ma < 3Mcr).

188
Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

Deflection or curvature has an inverse relationship to the second moment of area, so a


weighted average of flexibility and not stiffness should be used to compute deflection. Branson
(1963-68) initially expected a squared term in his expression for Ie in Equation (5.1). However
he found that the power in the (Mcr/Ma) term needed to be increased from 2 up to a value of 3 as
presented in Equation (5.2) for average member behaviour or 4 for section behaviour [73]. This
expression essentially represents a weighted average of the cracked (EcIcr) and uncracked (EcIg)
member stiffness [71]. Equation (5.3) was adopted by various codes of practice such as ACI-
435 (1966) and ACI-318 (1977) to compute for the effect of cracking in reduced stiffness of a
reinforced concrete section.

ͳ ‫ܯ‬ ͳ ‫ܯ‬ ͳ ͳ
ൌ ሺ ௖௥ൗ‫ ܯ‬ሻଷ ൅ ቀͳ െ ሺ ௖௥ൗ‫ ܯ‬ሻଷ ቁ ൒ ‫ݍܧ‬Ǥ ሺͷǤʹሻ
‫ܫ‬௘ ௔ ‫ܫ‬௚ ௔ ‫ܫ‬௖௥ ‫ܫ‬௚

‫ܯ‬ ‫ܯ‬
‫ܫ‬௘ ൌ ሺ ௖௥ൗ‫ܯ‬ሻସ ‫ܫ‬௚ ൅ ቀͳ െ ሺ ௖௥ൗ‫ܯ‬ሻସ ቁ ‫ܫ‬௖௥ ൑ ‫ܫ‬௚ ‫ݍܧ‬Ǥ ሺͷǤ͵ሻ

This method was intended to represent the sections at the working load and well below yield
of the reinforcement. Considering the RC elements subjected to only pure bending, this can be
regarded as a simple couple and neglecting the need to define a particular reference axis and
distinction between the centroid and the neutral axis, limits the application of Branson’s method
to model the tension stiffening [70]. Tension stiffening eventually becomes ineffective with
increasing the cracks, while in Branson’s model, the stiffness becomes asymptotic to the fully
cracked state but never reaches it [72].

Since some tension is transferred from steel to concrete through bond, so there are some
tensile stresses in the concrete between the cracks. Distribution and magnitude of the bond
stresses between the cracks represent the distribution of tensile stresses in the concrete and the
reinforcing steel between the cracks. This sharing of the tensile force between the reinforcement
and the concrete that represents difference between the actual and the zero tension response is
referred to “tension stiffening” [26],[37].

The average instantaneous moment-curvature response of RC elements under uniform


bending is shown as curve OAB in Figure (5.14). In reality, the flexural rigidity of the fully-
cracked cross-section (EcIcr) underestimates stiffness after cracking because of tension stiffening
that is represented by a reduction in average instantaneous curvature (δκ0.ts) as shown in Figure
(5.14) [74].

189
Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

Figure 5.14: Average moment versus instantaneous curvature relationship [74]

Bischoff (2004) [75] proposed the Equation (5.4) to model the gradual transition from the
gross (uncracked) moment of inertia (Ig) to the cracked transformed moment of inertia (Icr) as
loading progresses beyond the cracking moment Mcr.

‫ܫ‬௖௥
‫ܫ‬௘ ൌ ଶ ൑  ‫ܫ‬௚ ‫ݍܧ‬ǤሺͷǤͶሻ
‫ܯ‬
ͳ െ ߟ ቀ ௖௥ൗ‫ ܯ‬ቁ

ூ೎ೝ
Where; ߟ ൌ ͳ െ ூ೒
, Ma and Mcr are the service load moment and cracking moment of

section, Icr and Ig are cracking and gross moment of inertia of section.

Expression for Ie in Equation (5.4) was initially developed based on concepts of tension
stiffening to axial tension members (Bischoff and Paixao, (2004)) and later extended to include
flexural members (Bischoff, (2005)) resulting in the general model for Ie as shown in Equation
(5.5). This equation models the flexural stiffness EcIe corresponding to the Ma to define the
moment curvature response by introducing the tension stiffening factor β, varying between 1
(just before cracking with full tension stiffening) and 0 (no bond after cracking and no tension
stiffening) to account for tension carried by the concrete between cracks and stiffness of the
member cross-section [76].

‫ܫ‬௖௥
‫ܫ‬௘ ൌ ൑  ‫ܫ‬௚ ‫ݍܧ‬Ǥ ሺͷǤͷሻ
‫ܯ‬
ͳ െ ߟߚ ቀ ௖௥ൗ‫ ܯ‬ቁ

190
Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

As shown in Figure (5.15), setting β=1 representing the upper boundary on member
stiffness, gives a bilinear response with constant tension stiffening, while β=0 gives a lower
boundary with no tension stiffening that essentially gives the Icr response. Approximating β with
the ratio of Mcr / Ma gives a better model resulting to Equation (5.5) [75].

This tension-stiffening factor β, taken from the work of Rao (1966), and Leonhardt’s (1977)
group was instrumental in implementing the same tension-stiffening factor to compute crack
widths and estimate deflection in CEB-FIP (1978) for concrete structures, which forms the basis
of the later (CEB-1993) and Eurocode-2 (CEN-2004) [72].

As aptly shown in Figure (5.16), Branson’s equation overestimates the stiffness for flexural
members with Ig / Ic r > 3 or 4 and for 2 < Ig / Icr < 3, and the member response appears
reasonable. On the other hand, as tension stiffening far exceeds the upper limit given by β=1,
the computed member response is too stiff for a member with an Ig / Icr ratio of 15.

Figure 5.15: Theoretical moment-curvature response at the critical section using Bischoff’s method (Eq. 5.1)
with different tension stiffening values [76]

Figure 5.16: Deflection comparison using Branson’s (Eq. 5.1) and Bischoff’s expression for Ie [76]
191
Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

Bischoff (2005, 2007) [75] and Bischoff and Scanlon (2007) [77] modified the more realistic
and rational equation for Ie as a general expression in Equation (5.6), where β is used to account
for both shrinkage-induced cracking and the reduction in tension stiffening that occurs with
time. Early shrinkage in the days and weeks after casting will cause tension in the concrete and
a reduction in the cracking moment.

Confirming Bischoff (2005, 2007), the assumed value of Mcr which depends on the concrete
shrinkage prior to first loading, significantly affects the final predicted deflection [74]. The
effect of the service load moment to cracking moment (Ma / Mcr) on accuracy of the predicted
deflections is even more in lightly reinforced concrete and reinforce concrete members with
Fibre Reinforced Polymer (FRP) that typically has a service load of less than twice the cracking
value [72].

Scott and Beeby (2012) [78] showed that the tension stiffening and corresponding deflection
of a flexure member after cracking, are strongly influenced by the assumed value of the
cracking moment Mcr. The less assumed value of Mcr, the less stiffness and correspondingly
more deflection (see Figures (5.17) and (5.18)).

Computed deflection values of beams reinforced with steel at reinforcing ratios greater than
1%, using either equation (Branson’s or Bischoff’s) are similar, but for flexural members with
an Ig / Icr ratio greater than 3, differences become evident [1].

Due to the high sensitivity of the Ie value to the calculated value of Mcr, especially for lightly
loaded elements, failure to account for cracking due to unanticipated shrinkage restraint,
temperature gradients or construction loads can significantly underestimate the deflection, so the
upper limit of 0.6 Ig in Equation (5.6) is recommended by some researches. This limit has been
applied in AS-3600-09 [9]. The correct response depends mainly on the level of cracking and
lies somewhere between two extreme limits of the tension stiffening factor β as shown for
Equation (5.6) in Figure (5.18) [1].

‫ܫ‬௖௥
‫ܫ‬௘ ൌ  ൑ ‫ܫ‬௚ ‫ݍܧ‬Ǥ ሺͷǤ͸ሻ
‫ܯ‬௖௥ ‫ܫ‬
ͳ െ ߚሺ ሻሺͳ െ ௖௥ ሻ
‫ܯ‬௔ ‫ܫ‬௚

192
Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

Figure 5.17: Effect of tension stiffening on member stiffness [75]

Figure 5.18: Effect of cracking moment and tension stiffening on the member response after cracking [1]

5.11. EXISTING MODELS OF EFFECTIVE MOMENT OF INERTIA

The model which was initially developed by Branson (1963) was widely investigated by
other researchers worldwide. The basic concept of Branson’s model (effective moment of
inertia for a cracked section instead of gross moment of inertia) was modified by different
researchers and organizations mainly by implementing the idea of tension stiffening to apply to
different types of structures, loading rates, section reinforced with fibre and FRP sheets and
FRP bars. Equations (5.7) to (5.25) shown in Table (5.5) represent some verified models of
effective moment of inertia to enhance the prediction of short-term and long-term behaviour in
RC members. These models are developed to adapt to different load types and conditions,
reinforcing types, and supporting conditions. However, these models are generally applicable to
all types of loading and support conditions by modifying some additional factors.

193
Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

Table 5.5: Existing models of effective moment of inertia in RC section

Reference Equation Application Expression


 
ୣ ൌ ሺ ୡ୰ൗ ሻଶ ୥ ൅ ቀͳ െ ሺ ୡ୰ൗ ሻଶ ቁ ୡ୰ ൑
௔ ௔

୥ Eq. (5.7) Power in expressions;


Bronson (1963)   2: Initially expected
ୣ ൌ ሺ ୡ୰ൗ ሻଷ ୥ ൅ ቀͳ െ ሺ ୡ୰ൗ ሻଷ ቁ ୡ୰ ൑ Normal RC
Bronson (1965) ௔ ௔ 3: average member
୥ Eq. (5.8) beam
[71] behavior
 
ୣ ൌ ሺ ୡ୰ൗ ሻସ ୥ ൅ ቀͳ െ ሺ ୡ୰ൗ ሻସ ቁ ୡ୰ ൑ 4: section behavior
௔ ௔

୥ Eq. (5.9)

Al Shaik and Al ୣ ൌ ୡ୰ ൅ ൫ ୥ െ ୡ୰ ൯ሺ ୡ୰ൗ ሻଷି଴Ǥ଼஡ ൑ ୥ Point load on ρ: tensile reinforcement

Zaid (1993) [79] Eq. (5.10) beams ratio

ୣ ൑ ͲǤ͸ ୥ in AS3600-09
ACI-318-08 [34] ‫ܯ‬ ‫ܯ‬
‫ܫ‬௘ ൌ ሺ ௖௥ൗ ሻଷ ‫ܫ‬௚ ൅ ቀͳ െ ሺ ௖௥ൗ ሻଷ ቁ ‫ܫ‬௖௥ ൑ Normal proposed by Gilbert (2001)
ACI-435-68 [80] ௔ ௔

‫ܫ‬௚ Eq.(5.11) concrete for lightly reinforced


AS-3600-09 [9]
members
୑ౙ౨ ଷ ୑ౙ౨ ଷ
ୣ ൌ ቀ ቁ Ⱦୢ ୥ ൅ ൤ͳ െ ቀ ቁ ൨ ୡ୰ ൑ ୥ Ef: Elasticity of FRP bars
୑౗ ୑౗
ACI 440 (2006)
Eq. (5.12) FRP members Ef: Elasticity of steel bars
[81]
ா೑ ఘ
ߚௗ ൌ ߙ௕ ሺ ൅ ͳሻ, ߚௗ ൌ ͲǤʹ ൏ͳ
ாೞ ఘ್

Bischoff and
Originally for
Paixao (2004) [82] ‫ܫ‬௘௙ ൌ
ூ೎ೝ
൑  ‫ܫ‬௚ Eq. (5.13)

ଵିఎቀ ೎ೝൗெ ቁ
మ axial Applicable for flexural
Eurocode 2, ೌ
tension members
(1994) η = 1 - Icr/ Ig
members
[83]
Bischoff (2005,
୍ౙ౨
2007) [84] ୣ୤ ൌ ୑ ൑  ୥ Eq. (5.14)
ଵି஗ஒቀ ౙ౨ൗ୑ ቁ
౗ flexural
Bischoff and
no tension stiffening Ͳ ൑ Ⱦ ൑ ͳ full tension members
Scanlon (2007)
stiffening
[84]
ᇱ ୍ౙ౨ Simply
Bischoff and ୣ୤ ൌ ୑ ൑  ୥ Eq. (5.15) Equivalent moment of
ଵି஗ஓஒቀ ౙ౨ൗ୑ ቁ
౗ supported
Gross (2011) [72] inertia
γ=1.72 – 0.72Mcr /Ma , β= Mcr /Ma uniform load
Faza and
ଶଷூ೎ೝ ூ೐
GangaRao (1992) ‫ܫ‬௠ ൌ ଼ூ Eq. (5.16) FRP beam Ie from ACI-318
೎ೝ ାଵହூ೐
[85]
ெ೎ೝ ଷ ூ೒ ெ೎ೝ ଷ
Benmokrane et al. ‫ܫ‬௘ ൌ ቀ ቁ ൅ ͲǤͺͶ ൤ͳ െ ቀ ቁ ൨ ‫ܫ‬௖௥ ൑ ‫ܫ‬௚
ெೌ ଻ ெೌ FRP beam
(1996) [86]
Eq. (5.17)
Yost et al. (2003) ୑ౙ౨ ଷ ୑ౙ౨ ଷ
ୣ ൌ ቀ ቁ Ⱦୢ ୥ ൅ ൤ͳ െ ቀ ቁ ൨ ୡ୰ ൑ ୥
୑౗ ୑౗
[87] and ACI-440 FRP members

(2006) [81] Ⱦୢ ൌ ͲǤʹ ൏ͳ Eq. (5.18)
஡ౘ

194
Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

Table 5.5: Existing models of effective moment of inertia in RC section (conineud)

Reference Equation Application Expression


୑ౙ౨ ଷ ୑ౙ౨ ଷ ୍ౙ౨
ୣ ൌ ቀ ቁ Ⱦୢ ୥ ൅ ൤ͳ െ ቀ ቁ ൨ ൑ ୥
୑౗ ୑౗ ஓ
Rafi and Nadjai
Eq. (5.19) FRP members
(2009) [88]
ͲǤͲͲͳ͹ɏ ୤
ɀൌሺ ൅ ͲǤͺͷͶͳሻሺ ൅ ͳሻ
ɏୠ ʹୱ
୑ౙ౨
ୣ ൌ ୡ୰ ˆ‘” ൐͵ Eq. (5.20)
୑౗

Alsayed et al. ʹ ୡ୰


ୣ ൌ ൤ͳǤͶ െ ൬ ൰൨ ୡ୰ ൑ ୥ ˆ‘”ͳ FRP members
(2000) [89] ͳͷ ୟ
ୡ୰
൏ ൏͵
ୟ
ூ೒ ூ೎ೝ
ISIS Canada ‫ܫ‬௘ ൌ ಾ೎ೝ మ
൑  ‫ܫ‬௚
ூ೎ೝ ାሾଵି଴Ǥହቀ ቁ ሿ൫ூ೒ ିூ೎ೝ ൯
ಾೌ FRP members
(2001) [90]
Eq. (5.21)
β1=1:ribbed and 0:
ூ೒ ூ೎ೝ
Hall and Ghali ‫ܫ‬௘ ൌ ಾ೎ೝ మ ൑  ‫ܫ‬௚ smooth bars
ூ೎ೝ ାሾଵିఉభ ఉమ ቀ ቁ ሿ൫ூ೎ೝ ିூ೒ ൯
ಾೌ Normal concrete
(2000) [91] β2=1:first and 0:
Eq. (5.22)
sustained loading
୔ౙ౨ ଷ 
ୣ ൌ ቀ ቁ ୥ ൅ ቀͳ െ ሺ ୡ୰ൗሻଷ ቁ ୡ୰ ൑ ୥
ACI- 224.2R ୔ Members under Analogous approach for
-92 [92] Eq. (5.23) direct tension Ie in ACI-318
Ag: Gross cross section area , Acr: nAs
୑౗ ୐
ିሺ ൗ୑ ሻሺ ౙ౨ൗ୐ሻ஡
ୣ ൌ ୡ୰ ൅ ൫ ୥ െ ୡ୰ ൯ ౙ౨ ൑ ୥
Lcr : Cracked length
Fikry and Thomas for ρ>1% Eq. (5.24) FRP and Normal
L : Length of member
(1998) [93] ୑౗ ୐ concrete
ିሺ ൗ୑ ሻሺ ౙ౨ൗ୐ሻ
ୣ ൌ ୡ୰ ൅ ൫ ୥ െ ୡ୰ ൯ ౙ౨ ൑ ୥ ρ : Reinforcement ratio
Eq. (6.25)

5.12. SUMMARY

The second phase of the experimental program, i.e. the instantaneous and long-term
deflection under constantan loading and the load-deflection behaviour of slabs are presented in
Chapters Seven, Eight and Nine of this study.

In this chapter, a brief description of the serviceability aspects of the concrete structures
including the deflection limitations by the structure geometry is given. Additionally, some other
limitations and uncertainties in the literature and the reported experimental investigations about
inadequacy of the existing models to predict the flexural deflection of concrete members are
discussed.

195
Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

To calculate the long-term flexural deflection of reinforced concrete structures, majority of


the codes of practice imply the implication factor to the instantaneous deflection. A
comprehensive description of the ratio of long-term to instantaneous deflection as the multiplier
to estimate the time-dependent deflection is given. In addition, some existing experimental
results in the laboratory conditions and real in-situ data are analysed and discussed.

The effective parameters in the instantaneous and long-term deflection of reinforced concrete
structures are summarized. The tensile behavior of concrete and the interaction between the
concrete and reinforcement are among the most effective parameters in flexural deflection.
Therefore, a wide range of the existing models in the literature to describe the tensile behaviour
of concrete are presented and evaluated. The collected models developed since 1974 show
advances in describing the concrete behaviour under tension and the included parameters to
enhance the models.

The effective moment of inertia is another crucial parameter in studying the flexural
deflection of reinforced concrete structures. A background on the development of this concept
and the included parameters is given. Sensitivity of the effective moment of inertia to loading
value, size and ratio of reinforcement and the moment capacity of the reinforced concrete
section are also discussed. Existing models in the literature presented by the researchers and
codes of practice since 1963 are collected and evaluated.

196
Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

REFEERENCES

1. Gilbert, R.I., Creep and Shrinkage Induced Deflections in RC Beams and Slabs. ACI , Special
Publication, 2012. 284(13): p. 1-16.
2. 209.2R-08, A., Guide for Modeling and Calculating Shrinkage and Creep in hardened concrete.
American Concrete Institute (ACI), ACI Committee 209, Farmington Hills, MI 48331, U.S.A,
2008.
3. BS-5400-4, Steel, concrete and composite bridges, Part 4: Code of practice for design of
concrete bridges. Committee reference CSB/59, Draft for comment 88/11190 DC , British
Standard, 1999.
4. béton, C.e.-i.d., CEB-FIP model code 1990: design code. 1993: CEB Bulletin d’Information
No.213/214, Comité EuroInternational du Béton, Lausanne, Switzerland, Telford.
5. CSA.A23.3-04, Design of concrete structures. Canadian Standards Association (CSA), 2010.
6. EC-2, E., Design of Concrete Structures, Part 1: General Rules and Rules for buildings.
Commission of European Communities ENV, German version EN 1992-1-1:2004 + AC: 2010,
The Concrete Centre: Blackwater, Camberley, UK, . 2004: p. 1-1.
7. JSCE, Standard specification for concrete structures design. Japan society of Civil Engineering,
JSCE, Tokyo , Japan, 2007.
8. IS.456, Plain and Reinforced concrete- code of practice. Indian Standard, ICS, 91.100.30, New
Delhi, 2000.
9. AS-3600-09, Concrete structures. Australian Standard, 2009.
10. 1170.0, A.N., Structural design actions - General principles. Standards Australia, Sydney,
Australia, 2002.
11. Stewart, M.G., Serviceability reliability analysis of reinforced concrete structures. Journal of
Structural Engineering, 1996. 122(7): p. 794-803.
12. Branson, D.E., Deformation of concrete structures. 1977: McGraw-Hill Companies.
13. Jokinen, E.P. and A. Scanlon, Field measured two-way slab deflections. Canadian Journal of
Civil Engineering, 1987. 14(6): p. 807-819.
14. Unnikrishna Pillai, S. and D. Menon, Reinforced Concrete Design Tata Mc Graw Hill, New
Delhi, India,, 2003.
15. Nejadi, S., Time-dependent cracking and crack control in reinforced concrete structures. 2005,
PhD thesis, The University of New South Wales Sydney, Australia.
16. Taylor, P. and J. Heiman. Long-term deflection of reinforced concrete flat slabs and plates. in
ACI Journal Proceedings. 1977. ACI.
17. Gilbert, R., Shrinkage, cracking and deflection-the serviceability of concrete structures.
Electronic Journal of Structural Engineering, 2001. 1(1): p. 2-14.
18. Taha, M.R. and M.A. Hassanain, Estimating the error in calculated deflections of HPC slabs: a
parametric study using the theory of error propagation. ACI Special Publication, 2003. 210.
19. Ghali, A. and A. Azarnejad, Deflection prediction of members of any concrete strength. ACI
Structural Journal, 1999. 96(5).
20. Gilbert, R.I., Deflection Calculation for Reinforced Concrete Structures? Why We Sometimes
Get It Wrong. ACI Structural Journal, 1999. 96(6).
21. Heiman, J., A Comparison of Measured and Calculated Deflections of Flexural Memebers in
Four Reinforced Concrete Buildings. ACI Special Publication, 1974. 43.
22. Kaminetzky, D. and P. Stivaros, Construction Loads and Serviceability Requirements:
Deflection Control, Span/Thickness Limitations. ACI Special Publication, 2003. 210.
23. Vollum, R., R. Moss, and T. Hossain, Slab deflections in the Cardington in-situ concrete frame
building. Magazine of Concrete research, 2002. 54(1): p. 23-34.
197
Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

24. projec, E.c.b., Case studie on applying best practice to in-situ concrete frame buildings, St
George Wharf case study - slab deflection. European concrete building project, 2004.
25. French, C.E., et al. Field and Laboratory Study of Precast Composite Slab Span System
(PCSSS). in 2010 Concrete Bridge Conference: Achieving Safe, Smart & Sustainable Bridges.
2010.
26. Hirsch, J., et al., Practical Deflection Prediction of Concrete Slabs. ACI Special Publication,
2012. 284.
27. Uy, B. and M.A. Bradford, Ductility of profiled composite beams. Part I: Experimental study.
Journal of Structural Engineering, 1995. 121(5): p. 876-882.
28. Uy, B., Long-term service-load behaviour of simply supported profiled composite slabs.
Proceedings of the Institution of Civil Engineers: Structures and Buildings, 1997. 122(2): p. 193-
208.
29. BS-5950-82, tructural use of steelwork in building, Part 4. Code of practice for design of floors
with profiled steel sheeting. British Standard Institutation (BSI), London, BS 5950., 1982.
30. Gilbert, R. and X. Guo, Time-dependent deflection and deformation of reinforced concrete flat
slabs-an experimental study. ACI structural journal, 2005. 102(3).
31. Eigelaar, E.M., Deflections of reinforced concrete flat slabs. 2010, Stellenbosch: University of
Stellenbosch.
32. Park, H.-G., et al., Immediate and Long-Term Deflections of Reinforced Concrete Slabs Affected
by Early-Age Loading and Low Temperature. ACI Structural Journal, 2012. 109(3).
33. Lee, Y.H. and A. Scanlon, Comparison of one-and two-way slab minimum thickness provisions
in building codes and standards. ACI Structural Journal, 2010. 107(02).
34. ACI-318-02. Building code requirements for structural concrete (ACI 318-08) and commentary.
2002. American Concrete Institute, International Organization for Standardization.
35. BS-8110-1, Structural use of concrete, Part14: Code of practice for design and construction.
Committee reference B/525/2 Draft for comment 95/105430 DC, 1997.
36. Kaklauskas, G. and V. Gribniak, Eliminating shrinkage effect from moment curvature and
tension stiffening relationships of reinforced concrete members. Journal of Structural
Engineering, 2011. 137(12): p. 1460-1469.
37. Gilbert, R., The serviceability limit states in reinforced concrete design. Procedia Engineering,
2011. 14: p. 385-395.
38. Clarke, G., H. Scholz, and M. Alexander, New method to predict the creep deflection of cracked
reinforced concrete flexural members. ACI Materials Journal, 1988. 85(2).
39. Gilbert, R.I., Time effects in concrete structures. Developments in civil engineering, 1988. 23.
40. Murcia, J., Approximate time dependent analysis of reinforced concrete sections. Proposal of a
new factor for the calculation of long term deflections. Hormigón y Acero, 1991. 181: p. 9-17.
41. Favre, R. and H. Charif, Basic model and simplified calculations of deformations according to
the CEB-FIP model code 1990. ACI Structural Journal, 1994. 91(2).
42. Nie, J. and C.S. Cai, Deflection of cracked RC beams under sustained loading. Journal of
Structural Engineering, 2000. 126(6): p. 708-716.
43. Chiorino, M.A., A Rational Approach to the Analysis of Structural Effects due to Creep. ACI
Special Publication, 2005. 227.
44. Ghali, A., R. Favre, and M. Elbadry, Concrete structures: Stresses and deformations: Analysis
and design for serviceability. 2006: CRC Press.
45. Kang, Y.-J. and A.C. Scordelis, Nonlinear analysis of prestressed concrete frames. Journal of
the structural division, 1980. 106(2): p. 445-462.

198
Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

46. Mari, A.R., Nonlinear geometric, material and time dependent analysis of three dimensional
reinforced and prestressed concrete frames. 1984: Department of Civil Engineering, University
of California.
47. El-Badry, M.M. and A. Ghali, User's manual and computer program CPF: cracked plane
frames in prestressed concrete. 1985: University of Calgary. Department of Civil Engineering.
48. Ulm, F.-J., J.-L. Clément, and J. Guggenberger. Recent advances in 3D non-linear FE-analysis
of R/C and P/C Beam Structures. in Structures Congress XII. 1994. ASCE.
49. Damjanic, F. and D. Owen. Practical considerations for modelling of post-cracking concrete
behaviour for finite element analysis of reinforced concrete structures. in Proceedings,
International Conference on Computer Aided Analysis and Design of Concrete Structures, Split,
Yugoslavia, edited by Damjanic, F., Hinton, E., Owen, DRJ, Bicanic, N. and Simovic. 1984.
50. Torres, L., F. Lopez-Almansa, and L. Bozzo, Tension-stiffening model for cracked flexural
concrete members. Journal of Structural Engineering, 2004. 130(8): p. 1242-1251.
51. Scanlon, A. and D.W. Murray, Time-dependent reinforced concrete slab deflections. Journal of
the Structural Division, 1974. 100(9): p. 1911-1924.
52. Lin, C.-S. and A.C. Scordelis, Nonlinear analysis of RC shells of general form. Journal of the
Structural Division, 1975. 101(3): p. 523-538.
53. Vebo, A. and A. Ghali, Moment-curvature relation of reinforced concrete slabs. Journal of the
Structural Division, 1977. 103(3): p. 515-531.
54. Alih, S. and A. Khelil. Tension Stiffening Parameter in Composite Concrete Reinforced with
Inoxydable Steel: Laboratory and Finite Element Analysis. in Proceedings of World Academy of
Science, Engineering and Technology. 2012. World Academy of Science, Engineering and
Technology.
55. Scott, R. TECHNICAL NOTE. THE SHORT TERM MOMENT-CURVATURE RELATIONSHIP
FOR REINFORCED CONCRETE BEAMS. in ICE Proceedings. 1983. Thomas Telford.
56. Gilbert, R.I. and R.F. Warner, Tension stiffening in reinforced concrete slabs. Journal of the
structural division, 1978. 104(12): p. 1885-1900.
57. Schnobrich, W.C. The role of finite element analysis of reinforced concrete structures. in Finite
element analysis of reinforced concrete structures. 1985. ASCE.
58. Prakhya, G. and C. Morley, Tension-stiffening and moment-curvature relations of reinforced
concrete elements. ACI Structural Journal, 1990. 87(5).
59. Vecchio, F.J. and M.P. Collins. The modified compression-field theory for reinforced concrete
elements subjected to shear. in ACI Journal Proceedings. 1986. ACI.
60. Massicotte, B., A.E. Elwi, and J.G. MacGregor, Tension-stiffening model for planar reinforced
concrete members. Journal of Structural Engineering, 1990. 116(11): p. 3039-3058.
61. Kaklauskas, G., A New Stress-Strain Relationship for Cracked Tensile Concrete in Flexure.
Statyba, 1999. 5(6): p. 349-356.
62. Foster, S.J. and P. Marti, Cracked membrane model: Finite element implementation. Journal of
Structural Engineering, 2003. 129(9): p. 1155-1163.
63. Petersson, P.-E., Crack growth and development of fracture zones in plain concrete and similar
materials. 1981, Division, Inst.
64. Hillerborg, A., M. Modéer, and P.-E. Petersson, Analysis of crack formation and crack growth in
concrete by means of fracture mechanics and finite elements. Cement and concrete research,
1976. 6(6): p. 773-781.
65. Beeby, A., R. Scott, and A. Jones, Revised code provisions for long-term deflection calculations.
Proceedings of the ICE-Structures and Buildings, 2005. 158(1): p. 71-75.
66. Nayal, R. and H.A. Rasheed, Tension stiffening model for concrete beams reinforced with steel
and FRP bars. Journal of Materials in Civil Engineering, 2006. 18(6): p. 831-841.

199
Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

67. Stramandinoli, R.S. and H.L. La Rovere, An efficient tension-stiffening model for nonlinear
analysis of reinforced concrete members. Engineering Structures, 2008. 30(7): p. 2069-2080.
68. Ng, P.L., J.Y. Lam, and A.K. Kwan, Tension stiffening in concrete beams. Part 1: FE analysis.
Proceedings of the Institution of Civil Engineers: Structures and Buildings, 2010. 163(1): p. 19-
28.
69. Bacinskas, D., et al., Layer model for long-term deflection analysis of cracked reinforced
concrete bending members. Mechanics of Time-Dependent Materials, 2012. 16(2): p. 117-127.
70. Achintha, P.M. and C.J. Burgoyne, Moment-curvature and strain energy of beams with external
fiber-reinforced polymer reinforcement. ACI Structural Journal, 2009. 106(1).
71. Vakhshouri, B. and S. Nejadi. Limitations and Uncertainties in the Long Term Deflection
Calculation of Concrete Structures. in Second International Conference on Vulnerability and
Risk Analysis and Management (ICVRAM) and the Sixth International Symposium on
Uncertainty, Modeling, and Analysis (ISUMA). 2014.
72. Bischoff, P.H. and S.P. Gross, Equivalent moment of inertia based on integration of curvature.
Journal of Composites for Construction, 2010. 15(3): p. 263-273.
73. ACI-363-R, State of the art report on high-strength concrete. armington Hills (Michigan):
American Concrete Institute (ACI), 1992.
74. Bischoff, P.H. and M. Darabi, Unified approach for computing deflection of steel and FRP
reinforced concrete. ACI Special Publication, 2012. 284.
75. H, B.P., Discussion of “Tension Stiffening in Lightly Reinforced Concrete Slabs” R. Ian Gilbert ,
June 2007, Vol. 133, No. 6, pp. 899–903. Journal of structural engineering 2008. 134(7): p.
1259-1260.
76. Vollum, R.L., Influence of Construction Loading on Long-Term Slab Deflections. ACI Special
Publication, 2012. 284.
77. Bischoff, P.H. and A. Scanlon, Effective moment of inertia for calculating deflections of
concrete members containing steel reinforcement and fiber-reinforced polymer reinforcement.
ACI Structural Journal, 2007. 104(1).
78. Scott, R.H. and A.W. Beeby, Evaluation and management of tension stiffening. ACI Special
Publication, 2012. 284.
79. Alshaikh, A.H. and R. Al-Zaid, Effect of reinforcement ratio on the effective moment of inertia
of reinforced concrete beams. ACI Structural Journal, 1993. 90(2).
80. ACI-435, Deflections of reinforced concrete flexural member. American Concrete Institute
(ACI), Concrete International, 1968.
81. ACI-440-06, Guide for the design and construction of concrete reinforced with FRP bars.
American Concrete Institute (ACI) Committee 440, ACI 440.1R-06, ACI, Farmington Hills, MI,
44, 2006.
82. Bischoff, P.H. and R. Paixao, Tension stiffening and cracking of concrete reinforced with glass
fiber reinforced polymer (GFRP) bars. Canadian Journal of Civil Engineering, 2004. 31(4): p.
579-588.
83. du Béton, C.E.-I., CEB-FIP model code (MC-90). 1993, Thomas Telford Ltd, London.
84. Bischoff, P.H., Deflection calculation of FRP reinforced concrete beams based on modifications
to the existing Branson equation. Journal of Composites for Construction, 2007. 11(1): p. 4-14.
85. Faza, S. and H. GangaRao. Pre-and post-cracking deflection behaviour of concrete beams
reinforced with fibre-reinforced plastic rebars. in Proceedings of the First International
Conference on Advance Composite Materials in Bridges and Structures (ACMBS-I), Canadian
Society of Civil Engineers, Sherbrooke, Cananda. 1992.
86. Benmokrane, B., O. Chaallal, and R. Masmoudi, Flexural response of concrete beams reinforced
with FRP reinforcing bars. ACI Structural Journal, 1996. 93(1): p. 46-55.
87. Yost, J.R., S.P. Gross, and D.W. Dinehart, Effective moment of inertia for glass fiber-reinforced
polymer-reinforced concrete beams. ACI Structural Journal, 2003. 100(6).
200
Chapter 5: Mechanism of flexural deflection in reinforced concrete structures

88. Rafi, M.M., et al., Aspects of behaviour of CFRP reinforced concrete beams in bending.
Construction and Building Materials, 2008. 22(3): p. 277-285.
89. Alsayed, S., Y. Al-Salloum, and T. Almusallam, Performance of glass fiber reinforced plastic
bars as a reinforcing material for concrete structures. Composites Part B: Engineering, 2000.
31(6): p. 555-567.
90. Newhook J., Reinforcing concrete structures with fibre reinforced polymers. ISIS Canada:
Design Manual No.3, The Canadian Network of Centres of excellence on intelligent sensing for
innovative structures, Winnipeg, Manitoba, Canada, 2001.
91. Hall, T. and A. Ghali, Long-term deflection prediction of concrete members reinforced with
glass fibre reinforced polymer bars. Canadian Journal of Civil Engineering, 2000. 27(5): p. 890-
898.
92. ACI-224.2R-86, Cracking of concrete members in direct tension American Concrete Institute
(ACI) Committee 224, Detroit, USA, 1986.
93. Fikry, A.M. and C. Thomas, Development of a model for the effective moment of inertia of one-
way reinforced concrete elements. ACI Structural Journal, 1998. 95(4).

201
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

CHAPTER 6

EXPERIMENTAL PROGRAM (PHASE I) –


MATERIAL PROPERTIE OF
LIGHTWEIGHT CONCRETE CONTAINING
EXPANDED POLYSTYRENE BEADS
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

Table of Contents

6.1. INTRODUCTION ........................................................................................................... 202


6.2. LIGHTWEIGHT CONCRETE CONTAINING EXPANDED POLYSTRYRENE BEADS
204
6.2.1. Mixture Components .............................................................................................. 204
6.2.1.1. Natural aggregates .......................................................................................... 204
6.2.1.2. Coarse aggregate ............................................................................................ 204
6.2.1.3. Fine aggregates .............................................................................................. 206
6.2.1.4. Artificial lightweight aggregate ..................................................................... 207
6.2.1.5. Admixtures ..................................................................................................... 208
6.2.1.6. Cement ........................................................................................................... 209
6.3. MIXTURE PROPORTIONS ........................................................................................... 210
6.4. PREPARATION AND CURING CONDITION OF THE TEST SPECIMENS ............. 211
6.5. TEST METHODS OF THE HARDENED LIGHTWEIGHT CONCRETE ................... 211
6.6. PROPERTIES OF THE FRESH LIGHTWEIGHT CONCRETE ................................... 212
6.7. EXPERIMENTAL TEST RESULTS OF FRESH AND HARDENED LIGHTWEIGHT
CONCRETE ................................................................................................................................. 213
6.7.1. Slump and Fresh Density ........................................................................................ 213
6.7.2. Compressive Strength ............................................................................................. 213
6.7.2.1. Effects of shape and size of test specimens on the compressive strength ...... 215
6.7.3. Splitting Tensile Strength ....................................................................................... 216
6.7.3.1. Effect of shape and size of test specimens on splitting tensile strength ......... 218
6.7.4. Modulus of Elasticity .............................................................................................. 219
6.7.5. Modulus of Rupture (Flexural Tensile Strength) .................................................... 221
6.7.6. Drying Shrinkage .................................................................................................... 223
6.7.6.1. Curing and drying condition of shrinkage test at different ages ..................... 224
6.7.6.2. Proposed analytical models ............................................................................ 232
6.7.6.3. Rate of weight loss with time in shrinkage specimens ................................... 234
6.7.7. Creep Coefficient of Lightweight Concrete ............................................................ 234
6.7.7.1. Proposed analytical models ............................................................................ 239
6.7.7.2. Effect of shape and size of the test specimen on shrinkage strain .................. 242
6.7.7.3. Volume/surface ratio ...................................................................................... 243
6.7.8. Pull-Out Test ........................................................................................................... 243
6.7.8.1. Testing procedure ........................................................................................... 244
6.7.8.2. Loading rate ................................................................................................... 244
6.7.8.3. Load-displacement behaviour of the test specimens ...................................... 247
6.7.8.4. Effect of loading rate on the failure load ........................................................ 247
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

6.7.8.5. Bond-slip relationship .................................................................................... 248


6.7.9. Compressive Stress-Strain Curve of Lightweight Concrete Containing Expanded
Polystyrene Beads .................................................................................................................... 249
6.7.10. Energy Absorption under Compression .................................................................. 249
6.7.11. Mechanical Properties of Steel Bars ....................................................................... 253
6.7.12. Relationships between Compressive Strength and Mechanical Properties ............. 253
6.7.12.1. Splitting tensile strength ................................................................................. 254
6.7.12.2. Modulus of elasticity ...................................................................................... 255
6.7.12.3. Modulus of rupture ......................................................................................... 255
6.7.12.4. Ultimate pull-out load with respect to the compressive strength ................... 256
6.7.12.5. Compressive stress-strain relationship ........................................................... 257
6.8. COMPARING THE MECHANICAL PROPERTIES OF LIGHTWEIGHT, SELF-
COMPACTING AND CONVENTIONAL CONCRETE ............................................................ 259
6.8.1. Mixture Design ....................................................................................................... 259
6.8.2. Early-Age Mechanical Properties ........................................................................... 261
6.8.3. Time-Dependent Properties of Conventional, Self-Compacting and Lightweight
Concrete 264
6.8.3.1. Shrinkage strain.............................................................................................. 264
6.8.3.2. Creep strain .................................................................................................... 265
6.9. SUMMARY .................................................................................................................... 267
REFERENCES ............................................................................................................................. 280
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

List of Figures

Figure 6.1: BST aggregates in fresh lightweight concrete ............................................................ 208


Figure 6.2: Testing setup and different sizes and shapes of the specimens for compressive strength
tests ............................................................................................................................................... 214
Figure 6.3: Experimental vs. predicted values of compressive strength by Eq. (5.1) ................... 215
Figure 6.4: Different size and shapes of the compressive strength test specimens ....................... 215
Figure 6.5: Compressive strength of LWC by different size and shape of test specimens ........... 216
Figure 6.6: Specimen and test setup of splitting tensile strength .................................................. 217
Figure 6.7: Experimental vs. predicted values of splitting tensile strength by Eq. (5.2) .............. 218
Figure 6.8: Splitting tensile strength of cylindrical specimens ..................................................... 218
Figure 6.9: Specimen and test setup of modulus of elasticity ....................................................... 219
Figure 6.10: Development of modulus of elasticity of lightweight concrete with age ................. 220
Figure 6.11: Experimental vs. predicted values of modulus of elasticity by Eq. (6.3) ................. 220
Figure 6.12: Specimen and testing equipment of flexural tensile strength ................................... 221
Figure 6.13: Development of modulus of rupture of lightweight concrete by age ....................... 222
Figure 6.14: Experimental vs. predicted values of modulus of rupture by Eq. (5.4) .................... 222
Figure 6.15: Load-deflection in mid-span of the flexural test specimens at different ages .......... 223
Figure 6.16: Shrinkage specimens and testing equipment ............................................................ 224
Figure 6.17: Initial curing, drying and rewetting conditions of shrinkage prisms ........................ 224
Figure 6.18: Shrinkage variation in series A and C and shrinkage and swelling in series B ........ 226
Figure 6.19: Weight of shrinkage test prisms at different ages ..................................................... 227
Figure 6.20: Rate of weight loss of shrinkage specimens ............................................................. 228
Figure 6.21: Variation of weight loss rate with age of specimens ................................................ 229
Figure 6.22: Weight loss of specimens versus shrinkage strain .................................................... 230
Figure 6.23: Lost weight per initial weight versus shrinkage strain ............................................. 230
Figure 6.24: Rate of weight loss versus shrinkage rate ................................................................. 231
Figure 6.25: Rate of lost weight per initial weight versus shrinkage rate ..................................... 232
Figure 6.26: Experimental data of EPS-LWC versus predicted shrinkage strain by different models
...................................................................................................................................................... 233
Figure 6.27: Experimental data of EPS-LWC versus predicted shrinkage strain by different models
...................................................................................................................................................... 234
Figure 6.28: Creep test rig at two different loading levels ............................................................ 235
Figure 6.29: Creep coefficients of lightweight concrete under two different loading levels ........ 237
Figure 6.30: Comparing the ratio of creep coefficient at different ages ....................................... 237
Figure 6.31: Comparing the ratio of creep strain at different ages ............................................... 238
Figure 6.32: Development of creep and shrinkage strain of lightweight concrete by cylindrical
specimens...................................................................................................................................... 239
Figure 6.33: Comparison of the experimental and predicted creep strain at load level 1 ............. 240
Figure 6.34: Comparison of the experimental and predicted creep strain at load level 2 ............. 240
Figure 6.35: Comparison of the experimental and predicted creep coefficient at load level 1 ..... 241
Figure 6.36: Comparison of the experimental and predicted creep coefficient at load level 2 ..... 242
Figure 6.37: Shrinkage strain of EPS-LWC by prism and cylindrical specimen .......................... 243
Figure 6.38: Pull-out test specimens and testing equipment ......................................................... 244
Figure 6.39: Failure load of specimens in pull-out test at different ages (loading rate=72N/sec) 246
Figure 6.40: Brittle and ductile failure of pull-out test specimens at 28 and 56 days ................... 247
Figure 6.41: Comparison of failure load by different loading rates .............................................. 248
Figure 6.42: Stress-strain diagram of the LWC mixture at different ages .................................... 249
Figure 6.43: The energy stored under compression at different ages ........................................... 250
Figure 6.44: Predicted versus experimental energy absorption under compression ..................... 251
Figure 6.45: Stored energy under compression versus strain of EPS-LWC at different ages ....... 252
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

Figure 6.46: Compressive stress-strain variation of the LWC mixture in elastic range at different
ages ............................................................................................................................................... 252
Figure 6.47: Tensile Stress- strain curve of the N12 steel bars ..................................................... 253
Figure 6.48: Predicted versus experimental splitting tensile strength of EPS-LWC .................... 254
Figure 6.49: Predicted versus experimental modulus of elasticity of EPS-LWC ......................... 255
Figure 6.50: Predicted versus experimental modulus of rupture of EPS-LWC ............................ 256
Figure 6.51: Ascending ultimate pull-out load by age development of compressive ................... 257
Figure 6.52: Comparison of the measured and predicted stress-strain variation under compression
...................................................................................................................................................... 258
Figure 6.53: Compressive strength of EPS-LWC, SCC and CC mixtures at 14 and 28 days ...... 262
Figure 6.54: Modulus of elasticity of EPS-LWC, SCC and CC mixtures at 14 and 28 days........ 262
Figure 6.55: Increasing rate of compressive strength in EPS-LWC, SCC and CC mixtures ........ 262
Figure 6.56: Increasing rate of modulus of elasticity in EPS-LWC, SCC and CC mixtures ........ 263
Figure 6.57: Comparison of the modulus of rupture in LWC, CC and SCC mixtures ................. 264
Figure 6.58: Comparison of shrinkage strain in EPS-LWC, SCC and CC mixtures .................... 265
Figure 6.59: Creep coefficient of CC, SCC and EPS-LWC specimens loaded at age 14 days ..... 266
Figure 6.60: Initial curing, drying and rewetting conditions of shrinkage prisms ........................ 270
Figure 6.61: Experimental data of EPS-LWC versus predicted shrinkage strain by different models
...................................................................................................................................................... 270
Figure 6.62: Creep coefficients of LWC specimens under sustained loads of 30% and 40% of the
ultimate capacity ........................................................................................................................... 271
Figure 6.63: Comparing the decreasing ratio of creep coefficient at different ages ..................... 272
Figure 6.64: Stress-strain diagram of the LWC mixture at different ages .................................... 274
Figure 6.65: Stored energy under compression versus strain of EPS-LWC at different ages ....... 275
Figure 6.66: Compressive strength of EPS-LWC, SCC and CC mixtures at 14 and 28 days ....... 277
Figure 6.67: Modulus of elasticity of EPS-LWC,SCC and CC mixtures at 14 and 28 days........ 277
Figure 6.68: Comparison of the modulus of rupture in LWC, CC and SCC mixtures ................. 278
Figure 69: Comparison of shrinkage strain in EPS-LWC, SCC and CC mixtures ....................... 278
Figure 6.70: Creep coefficient of CC, SCC and EPS-LWC specimens loaded at age 14 days ..... 279
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

List of Tables

Table 6.1: Conducted tests on lightweight, self-compacting and conventional concrete specimens
...................................................................................................................................................... 203
Table 6.2: Properties of Dunmore 10 mm coarse aggregate ......................................................... 205
Table 6.3: Properties of Peppertree-P coarse sand ........................................................................ 206
Table 6.4: Properties of washed Kurnell fine sand ....................................................................... 207
Table 6.5: General information about the used water reducing admixture (Plastiment-20) ......... 209
Table 6.6: Properties of the Shrinkage Limited (SL) type cement ................................................ 210
Table 6.7: Mix design of lightweight concrete containing expanded polystyrene beads (SSD
condition) ...................................................................................................................................... 211
Table 6.8: Shape and size of the LWC test specimens and test ages ............................................ 212
Table 6.9: Applied codes on the curing, sampling and testing the LWC specimens .................... 212
Table 6.10: Test methods to determine the fresh properties of LWC ........................................... 213
Table 6.11: Compressive strength of lightweight concrete mixture at different ages ................... 214
Table 6.12: Splitting tensile strength of LWC mixture at different ages ...................................... 217
Table 6.13: Modulus of elasticity of LWC mixture at different ages ........................................... 219
Table 6.14: Modulus of rupture of lightweight concrete at different ages .................................... 221
Table 6.15: Recorded shrinkage strain and weight of specimens at different ages ....................... 225
Table 6.16: Creep strain and creep coefficient of lightweight concrete cylindrical specimen (load
level 1) .......................................................................................................................................... 236
Table 6.17: Creep strain and creep coefficient of lightweight concrete cylindrical specimen ( load
level 2) .......................................................................................................................................... 236
Table 6.18: Shrinkage strain of cylindrical specimens (companion of creep specimens) ............. 238
Table 6.19: Coefficients of creep strain of EPS-LWC in Eq. (6.7)............................................... 240
Table 6.20: Coefficients of creep strain of EPS-LWC in Eq. (6.7)............................................... 241
Table 6.21: Specimen dimensions and variables of pull-out test .................................................. 246
Table 6.22: Bond stress and corresponding slip values of test specimens at different ages ......... 248
Table 6.23: The energy absorbed in cylindrical specimens under compression ........................... 250
Table 6.24: Mechanical properties of N12 steel bar ..................................................................... 253
Table 6.25: Mechanical properties of EPS-LWC at different ages ............................................... 254
Table 6.26: Correlation factor between the predicted and experimental mechanical properties .. 256
Table 6.27: Mixture design of self-compacting concrete [1] and lightweight concrete ................ 260
Table 6.28: Physical and mechanical properties of fibres in self-compacting concrete mixture [1]
...................................................................................................................................................... 260
Table 6.29: Mechanical properties of self-compacting concrete [1], lightweight concrete and
conventional concrete [2] mixtures ............................................................................................... 261
Table 6.30: Free shrinkage strain of CC, LWC and SCC mixtures .............................................. 264
Table 6.31: Creep coefficient of CC, LWC and SCC mixtures after 240 days sustained loading 266
Table 6.32: Shape and size of the LWC test specimens and test ages .......................................... 267
Table 6.33:Applied codes on the curing, sampling and testing the LWC specimens ................... 268
Table 6.34:Test methods to determine the fresh properties of lightweight oconcrete .................. 268
Table 6.35: Mix design of lightweight concrete containing expanded polystyrene beads (SSD
condition) ...................................................................................................................................... 268
Table 6.36: Mechanical properties of EPS-LWC at different ages ............................................... 269
Table 6.37: Bond stress and corresponding slip values of test specimens at different ................. 273
Table 6.38: The energy absorbed in cylinder specimens under compression ............................... 274
Table 6.39: Mixture design of self-compacting and lightweight concrete .................................... 276
Table 6.40: Mechanical properties of SCC, LWC and CC mixtures ............................................ 276
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

Chapter 6: Material properties of lightweight concrete containing


expanded polystyrene beads

6.1. INTRODUCTION

This chapter explains the experimental program investigating the mechanical


characteristics of the Light-Weight Concrete (LWC) in fresh and hardened states. The
experimental program has been conducted in the concrete laboratory of University of
Technology Sydney (UTS).

The main objective of this study is to compare the short-term (instantaneous) and long-
term deflection of one-way slabs made with different types of concrete, i.e. Conventional
Concrete (CC) and Self-Compacting Concrete (SCC) from previously conducted
researches and Light-Weight Concrete (LWC) in the present study. In this regard, the
experimental results of the fresh and hardened properties of SCC conducted by Aslani
(2014) [1] and CC carried out by Nejadi (2005) [2] have also been compared with the
corresponding properties of LWC in this study. Comparative analysis of the material
properties in CC, SCC and LWC and development of the mechanical properties with age
are the major objectives of this section.

In terms of the fresh concrete properties, the fresh density and slump value from the
mixture design of CC, SCC and LWC are investigated. In total, one mixture for LWC,
one mixture for CC and four mixtures for SCC have been considered in evaluating the
fresh and hardened properties of concrete. The mixtures for SCC include one plain
mixture and three fibre reinforced mixtures of steel, polypropylene and hybrid (steel and
polypropylene) fibres mixed with different proportions.

In a hardened state, the compared properties of CC, LWC and SCC include the
Compressive Strength (CS), Modulus of Elasticity (MoE), and Modulus of Rupture
(MoR), Splitting Tensile Strength (STS), and bond from pull-out test, and Compressive
202
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

Stress-Strain Curve (CSSC). Also, the time-dependent properties like creep and shrinkage
strains from these types of concrete are investigated.

To determine the mechanical properties of CC SCC and LWC, different types of tests
on the concrete specimens have been conducted. Table (6.1) shows the types of tests
performed to obtain the concrete properties used in the experimental and numerical
investigations of the slabs in Chapters Six, Seven, Eight and Nine of this study.

Table 6.1: Conducted tests on lightweight, self-compacting and conventional concrete specimens

Hardened concrete tests Fresh concrete tests


Concrete type

Compressive

Stress-strain
Modulus of

Shrinkage
Elasticity
Modulus

Splitting

Pull-out
strength

strength

Density
rupture

Slump
tensile

Creep
CC * * * * * * * * * *

SCC * * * * * * * * *

LWC * * * * * * * * * *

Other tests by Aslani (2014) [1] such as slump flow, T50cm time, J-ring flow, V-
funnel flow time, and L-box blocking ratio tests are applicable just for SCC in the fresh
state and are not comparable with the fresh properties of LWC and CC in this study.
These tests for SCC are performed to check the filling ability, passing ability and
segregation resistance of SCC mixtures.

In addition to the comparable experimental results of CC and SCC properties, to better


understand the post-cracking behaviour of LWC, flexural toughness characteristics
analysis and energy absorption in compression has been performed in this study also.

The main parts of this chapter can be summarized as follows:

- Description of the mixture proportions of LWC along with the chemical, physical
and mechanical properties of the components;
- Evaluation and analysis of the experimental results of the tests on fresh LWC
mixture and the hardened LWC specimens;
- Proposing the models to predict the mechanical properties of LWC;
- Comparison of the corresponding material or mechanical characteristics of LWC
in this study with SCC and CC from previously conducted studies.

203
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

6.2. LIGHTWEIGHT CONCRETE CONTAINING EXPANDED


POLYSTRYRENE BEADS

6.2.1. Mixture Components


The LWC used in this study contains Expanded Poly-Styrene (EPS) beads as the
lightweight aggregates to reduce the mixture density. LWC containing EPS beads (EPS-
LWC) is widely used in Australia. It is also a common construction material in floating
decks worldwide; however, it has not been investigated as much as the other types of
LWC and CC. In the mechanical properties and time-dependent behaviour of the test
specimens and structures made with EPS-LWC reported in the literature have been
collected and presented in Chapter Three.

Considering the limited investigations of the mechanical properties of EPS-LWC, the


existing correlations between the experimental and predicted data are unclear; therefore
regression analyses are conducted on existing experimental data to propose new models
of predicting its mechanical properties in Chapter Three. In this Chapter, components of
the utilized LWC and their proportions in the mixture are included. The mixture design of
EPS-LWC in this study contains natural coarse and fine aggregates, artificial lightweight
aggregate, cement, water and admixture. Each component is described as follows:

6.2.1.1. Natural aggregates

As discussed in Chapter Two, the main difference between the mixture design of CC
and LWC is the replacement of some or whole part of the normal weight (natural) coarse
aggregate with lightweight (natural or artificial) aggregates in LWC. The mixture design
of LWC in this study contains both types of natural coarse aggregates and artificial
lightweight aggregate. Moreover, a blend of two types of natural fine aggregates has been
used in the EPS-LWC mixture.

6.2.1.2. Coarse aggregate

The crushed Latite aggregate with a maximum size of 10 mm from Dunmore quarry
(NSW, Australia) is used as the natural coarse aggregate. The properties of Dunmore
coarse aggregate, including the chemical components and physical and mechanical
properties are presented in Table (6.2). Australian standard, AS-1141 (2011) [3] along
with the Regional Transportation Authority (RTA) (2006) and Road and Maritime
services Specifications (RMS) (2015) recommendations [4] were applied in sampling and
testing of the aggregates. The applied standards and the minimum or maximum allowable

204
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

values in each case are also shown in the Table (6.2). The measured properties of
Dunmore coarse aggregate are in the allowable range.

Table 6.2: Properties of Dunmore 10 mm coarse aggregate

Test method Test Allowable Measured


Passing A.S. sieve (%)
13.2 mm 100 100
9.5 mm 85-100 88
AS-1141.11.1 (2009) 53
6.7 mm 40-60
[5]
4.75 mm 0-20 20
2.36 mm 0-5 4
1.18 mm 2
AS-1141.12 (2011) [6] ≤ 75 μm (%) 1
Mis-shaped particles (%)
AS-1141.14 (2011) [7] Ratio 2:1 <25 23
Ratio 3:1 <10 6
Average least dimension (mm) 3.5
RMS-T235 Fraction (-9.5+6.7mm) 6.1
Fraction (-6.7+4.75mm) 4.4
Aggregate shape by the ratio of greatest to least
di
0-2 mmi 5.6
2-4 mm 2.5
RMS-T278 [4] 4-6 mm 2.3
6-8 mm 2.0
The ratio of AGD to ALD 2.49
Crushed particles derived from gravel
AS-1141.18 (2011) [8] % of crushed particles ≥80 100
% of uncrushed particles Nil
Uncompacted bulk density (t/m3) 1.33
AS-1141.4 (2011) [9] Compacted bulk density (t/m3) ≥1.2 1.5
RMS-T262 Moisture content (%) 3.0
Particle density (dry) (t/m3) ≥2.1 2.63
AS-1141.6.1 (2011) Particle density (SSD) (t/m3) 2.66
[10] Apparent Particle density (t/m3) 2.76
Water absorption <2.5 1.9
Average aggregate dry strength (KN) 366
Average aggregate wet strength (KN) ≥100 246
AS-1141.22 (2008) Wet/dry strength var. (%) <35 33
[11]
Fraction tested (mm) -9.5+6.7
The amount of significant breakdown (%) <0.2
AS-1141.23 (2009) Los Angeles value grading “K” (% loss) < 30 15
Sodium sulphate soundness (Total weighted % loss) < 6.0 1.0
Fraction tested
AS-1141.24 (2013) -13.2mm + 9.5mm (% loss) 0.6
[13]
-9.5mm + 4.75mm (% loss) 0.7
-4.75mm + 2.36mm (% loss) 2.6
Weak particles (%)
AS-1141.32 (2008)
[14] % of original sample passing 2.36 mm (sieve 3.4) < 0.5 Nil
Degradation factor- coarse aggregate 66
As-1141.25.2 (2013)
[15] The washed water after using permitted 500 ml Not clear
SSD : Saturated Surface-Dry

205
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

6.2.1.3. Fine aggregates


In this type of LWC, a blend of 50% coarse sand and 50% fine sand provides optimum
mix. An equal weight proportion of Peppertree-P sand with a maximum size of 5 mm,
and washed Kurnell Natural River sand with a maximum size of 2 mm were used as fine
aggregates in the mixture. Australian standard AS-1141(2011) [3] along with the
Regional Transportation Authority (RTA) (2006) and RMS (2015) codes [4] were applied
in sampling and testing of the aggregates. The properties of Peppertree-P coarse sand and
washed Kurnell Natural River sand are presented in Tables (6.3) and (6.4) respectively.
Testing and sampling standards are also illustrated in those Tables.

Table 6.3: Properties of Peppertree-P coarse sand


Test method Test Allowable Measured
Passing A.S. sieve (%)
6.7 mm 100 100
4.75 mm 95-100 100
2.36 mm 73-88 80
AS-1141.11.1(2009)
1.18 mm 55-70 55
[5]
600 μm 35-50 37
425 μm 30
300 μm 25-35 23
150 μm 10-20 11
AS-1289.4.3.1. PH value of soil 8.8
Chloride as Cl (%) 0.007
AS-1012.20
Sulphate as SO3 0.03
AS-1141.12 (2011) ≤ 75 μm (%) 5-10 5
AS-1141.13 (2007) ≤ 2 μm (%) 1 0.7
Uncompacted bulk density (t/m3) 1.69
AS-1141.4 (2011) [9]
Compacted bulk density (t/m3) ≥1.2 1.86
RMS-T262 Moisture content (%) 0.3
Particle density (dry) (t/m3) ≥2.1 2.69
AS-1141.6.1(2011) Particle density (SSD) (t/m3) 2.71
[10] Apparent Particle density (t/m3) 2.76
Water absorption <2.0 0.9
Sodium sulphate soundness (Total weighted % loss) < 5.0 0.9
Fraction tested
AS-1141.24(2013)
-4.75 mm + 2.36 mm (% loss) 1.1
[13]
-2.36 mm +1.18 mm (% loss) 0.8
-1.18 mm + 600 μm (% loss) 0.7
-600 μm + 300 μm (% loss) 0.9
Weak particles (%)
AS-1141.32 (2008)
[14] The % of original sample passing 2.36 mm (sieve 3.4) < 0.5 0.9
As-1141.25.2(2013) Degradation factor- fine aggregate ≥ 60 85
[15] The washed water after using permitted 500 ml clear
Organic impurities other than sugar
AS-1141.34(2007) Not darker 0.2 % ,
[17] The colour assessment was made visually using than std. Pass
coloured reference glass
AS-1141.35(2007) Sugar Negative Not detected
Method of determining voids content
RMS-T279 % of voids 38-44 40.7
The mean flow time (sec.) 21-27 27.3
Micro-deval abrasion test
ASTM-D7428 % loss 15.3
The % loss of the control aggregate tested closest to 19.5
AS-1289.3.7.1 Sand equivalent of the soil using a power-operated ≥ 60 76
RMS-T659 Average Methylene blue adsorption value (mg/g) 5 3

206
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

Table 6.4: Properties of washed Kurnell fine sand

Test method Test Allowable Measured


Passing A.S. sieve (%)
6.7 mm 100 100
4.75 mm 90-100 100
2.36 mm 60-100 100
AS-1141.11.1
1.18 mm 30-100 100
[5]
600 μm 15-100 99
425 μm 91
300 μm 5-50 58
150 μm 0-20 2
AS-1289.4.3.1. PH value of soil 8.8
Chloride as Cl (%) 0.003
AS-1012.20
Sulphate as SO3 0.09
AS-1141.12 ≤ 75 μm (%) 0-5 Nil
Not
AS-1141.13 ≤ 2 μm (%)
applicable
AS-1141.31 Light particles (%) Nil
3
Uncompacted bulk density (t/m ) 1.33
AS-1141.4 3
Compacted bulk density (t/m ) ≥1.2 1.47
RMS-T262 Moisture content (%)
Particle density (dry) (t/m3) ≥2.1 2.52

AS-1141.6.1 Particle density (SSD) (t/m3) 2.55


[10] Apparent Particle density (t/m ) 3
2.59
Water absorption 1.0
Sodium sulphate soundness < 5.0
AS-1141.24 [13] Total weighted (% loss) C,B2<12 0.9
Fraction tested (-600 μm +300 μm (% loss) A1,A2,B1<15 0.9
Organic impurities other than sugar
Not darker than
AS-1141.34 [17] The colour assessment was made visually using Pass
std.
coloured reference glass
Not
AS-1141.35 [18] Sugar Negative
detected
AS-1141.33 Silt content (%) 3
SSD : Saturated surface-dry

6.2.1.4. Artificial lightweight aggregate

The low-density foam of Expanded Polystyrene (EPS) is used as lightweight aggregate


to reduce the density of concrete. EPS is a closed-cell lightweight cellular plastic material
produced from polystyrene. The material has been modified by the addition of flame

207
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

retardant additives. Polystyrene is literally translated as “polymerized styrene”. It is the


single styrene molecules that are chemically joined to form a large molecule which is
called the polymer. Styrene is produced from benzene and ethylene, and polymerization
is accomplished in the presence of catalysts, usually organic peroxides. The expandable
foam is produced as small beads.

“BST” aggregate is a commercial name for the used spherical-shaped polystyrene


beads with hydrophilic type chemical coating. The percentage passing of the BST beads
on 4.75 mm and 2.36 mm sieves are 100% and 90% respectively. Bulk and particle
density of the beads are 35 and 67 kg/m3 respectively. Figure (6.1) shows the BST
aggregates and the fresh LWC concrete mixture containing the BST beads utilized in this
study. The mechanical and chemical properties of EPS beads are comprehensively
discussed in Chapter Three.

Figure 6.1: BST aggregates in fresh lightweight concrete

6.2.1.5. Admixtures

The concrete containing BST aggregates (spherical-shaped polystyrene beads) is


commercially known as BST concrete. Water-reducing admixtures are used with BST
concrete in general. Plastiment-20 in agreement with AS-1478.1 (2000) [19] is used as
water reducing agent in the LWC mixture in this study. It is based on lignosulfonates.
This admixture helps to achieve increased workability without loss of strength and also to
increase strength without loss of workability. The time period during which the concrete
can be placed is also extended. Melamine based products do not perform well with LWC
containing BST. Sika-Australia supplies Plastiment-20 as a water reducing admixture in
the concrete. Table (6.5) presents general information about the applications, advantages
and storage conditions of this product.

208
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

Table 6.5: General information about the used water reducing admixture (Plastiment-20)

Description Water reducer admixture based on selected lignosulfonates

- To provide a cost effective solution in enhancing the water reduction


- To improve the efficiency of the water content in concrete mixes
Application
- To achieve increased workability without loss of strength
- To increase strength without loss of workability
- Improved Workability and Finish-ability
- Higher slump can be achieved without further addition of water
- Easier and quicker placing and compaction
- Excellent Strength Performance
- Enhances the strength gained at all ages
Advantages
- Homogeneous concrete with less shrinkage requiring lower pumping pressure
- Improve plastic concrete properties
- Improves cohesiveness with improved placement properties and minimises the
segregation and bleeding problems.
- Retardation without loss in strength
- Between 0°C and 35°C in unopened original containers

Storage and shelf-life - Protected from direct sunlight and frost


- Shelf-life is at least one year.

- Added with the initial batching water. Never added to dry mix
- Recommended dosage is 350 ml ± 100 ml per 100 kg of total cementitious
How to use
material, Optimum dosage should be determined by site trials with the
particular concrete mix design

6.2.1.6. Cement

BST concrete mix design is based on the Shrinkage Limited (SL) cement confirming
AS-3972 (2010) [20]. This cement is a special and modified type of General Purpose
(GP) cement. Other types of cement are suitable, but their effect on yield and
performance of the mixture should be determined. The chemical, physical and
mechanical properties of the used SL (GP) type cement are presented in Table (6.6).
These properties are in agreement with the specifications defined in AS-2350 (2006) [21].
It is worth mentioning that the used SL type cement contains up to 7.5% additional
limestone mineral powder.

209
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

Table 6.6: Properties of the Shrinkage Limited (SL) type cement

Test method Allowable Measured


Chemical components
LOI (%) 4.3 4.1
SO3 (%) (AS-2350.2) < 3.5 3.1
CI (%) (AS-2350.2) < 0.10 0.02
CaO (%) 64.3 62.6
SiO2 (%) 19.4 19.26
Al2O3 (%) 5.15 5.15
Fe2O3 (%) 3.18 3.08
AS-2350.2 (2006) [22] MgO (%) 1.59 1.14
K2O (%) 0.51 0.53
Na2O (%) 0.18 0.08
P2O5 (%) 0.1 0.05
Mn2O3 (%) 0.08 0.06
Cr2O3 (%) 0.01 0.01
SrO (%) 0.05 0.04
NaEq (%) 0.4 0.3
Physical properties
AS-2350.8(2006) [23] Fineness index (m2 /kg) 441 395
AS-2350.3 (2006) [24] Normal consistency (%) 27.7
Mechanical properties
AS-2350.13 (2006) [25] Residue (45 μm ) 2.4
Drying shrinkage - 28 days mean (με) < 750 570
AS-2350.11 (2006) [26] Compressive strength
݂௖ᇱ - 3 days (MPa) 34
݂௖ᇱ - 7 days (MPa) 35 45.6
݂௖ᇱ - 28 days (MPa) 45 61.1
AS-2350.4 (2006) [27] Setting time
Initial (minute) ≥ 45 105
Final (minute) < 600 150
AS-2350.5 (2006) [28] Soundness <5 mm 1mm

6.3. MIXTURE PROPORTIONS

The consistent materials and corresponding mixture proportions that used to achieve a
structural concrete with low density in the range of 2000 kg/m3 is presented in Table
(6.7). The quantities are given for each cubic metre of the LWC mixture.

210
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

Table 6.7: Mix design of lightweight concrete containing expanded polystyrene beads (SSD condition)

Component Proportion
3
Cement (kg/m ) 500
3
Water (litre/ m ) 180
Water to cement ratio (w/c) 0.36
BST aggregate (Litre/ m3) 300
3
Coarse sand (kg/ m ) 310
Fine aggregate 3
Fine sand (kg/ m ) 310
3
Coarse aggregate (kg/ m ) 800
Admixture (litre/ m3) 2
3
Mixture volume (m ) 1.3
Slump (mm) 130
3
Fresh density (kg/m ) 2000

6.4. PREPARATION AND CURING CONDITION OF THE TEST


SPECIMENS

All specimens were cast within 15 minutes after the end of the mixing process. The
fresh concrete was poured into the moulds in 2 or 3 layers following the seconds of
vibration by the vibrating table. Vibrating the moulds over 5 seconds causes the EPS
beads to float up to the concrete surface. After casting, specimens were carefully moved
and kept covered in a controlled chamber at 23 ±2 oC conforming AS-1210.8.1 [29] up to
the demoulding time after 24 hours. Then, all test specimens were stored in water pre-
saturated with lime until the measurement time. Test ages are illustrated in Table (6.8).
All tests were performed using different sizes of the cylinder, cube and prism specimens.
For further investigation of the effects of curing conditions on shrinkage behaviour,
various curing conditions were applied to the shrinkage prisms that will be explained in
the following sections.

6.5. TEST METHODS OF THE HARDENED LIGHTWEIGHT


CONCRETE

The test ages, shape and size of the specimens for hardened LWC are shown in
Table (6.8). Three LWC specimens have been tested per each age and property.

211
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

Table 6.8: Shape and size of the LWC test specimens and test ages

Cylinder Cube Prism Test ages


Property
(mm) (mm) (mm) (days)
Compressive strength 75×150, 100×200,150×300 100, 150 3,7, 14, 21, 28, 56, 91
Modulus of elasticity 150×300 3,7, 14, 21, 28, 56, 91
Modulus of rupture 100×100×400 3,7, 14, 21, 28, 56, 91
Splitting tensile
100×200,150×300 3,7, 14, 21, 28, 56, 91
strength
Steel-concrete Bond 150×300 3,7, 14, 21, 28, 56, 91
Pull-out test 150×300 7, 14, 28, 56, 91
Stress-strain 3,7, 14, 21, 28, 56, 91
Creep 150×300 Up to 210 days
Shrinkage 75×75×280 Up to 450 days

Australian codes and some international codes are applied to determine the mechanical
properties of LWC in this study. Table (6.9) explains the applied Australian standards
and international codes for each type of the test under laboratory conditions.

Table 6.9: Applied codes on the curing, sampling and testing the LWC specimens

Property Test method Property Test method

AS 1012.8.1 (2014) , AS 1012.8.2 (2014),


Compressive strength Modulus of rupture
ASTAM C39 (2000) ASTM C1018 (2002)

AS 1012.8.1(2014),
Splitting Tensile strength Steel-concrete Bond RILEM –RC6 (2004)
ASTM C496 (2000)

Creep AS 1012.16 (2014) Pull-out test RILEM –RC6 (2004)

AS 1012.17 (2014),
Shrinkage AS-1012.13 (2014) Modulus of elasticity
ASTM C469 (2002)

6.6. PROPERTIES OF THE FRESH LIGHTWEIGHT CONCRETE

The required measurements of LWC in a fresh state are generally similar to those of
CC, however it is almost entirely different from the fresh state of SCC. The specific
gravity and slump values are the measured parameters of fresh LWC in this study. The
experimental tests required for the LWC are implemented worldwide under laboratory
conditions. These experimental investigations evaluate the density and workability of
212
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

fresh LWC. Table (6.10) explains the Australian standards AS-1012 (2014) [30] applied
to curing, sampling, slump test and mass per volume measurements of the fresh LWC.

Table 6.10: Test methods to determine the fresh properties of LWC

Curing Sampling Slump test Mass per volume

Test method AS-1012.8 (2014) AS-1012.1 (2014) AS1012.3.1 (2014) AS1012.5 (2014)

6.7. EXPERIMENTAL TEST RESULTS OF FRESH AND HARDENED


LIGHTWEIGHT CONCRETE

6.7.1. Slump and Fresh Density


According to test results, the slump value for the LWC was 130 mm. This slump value
provided the required flow-ability and workability to pour fresh concrete in the test
specimens and the reinforced concrete slabs. Density of the concrete in the fresh state was
2000 kg/m3. Additionally, to avoid any blocking problem, the maximum size of the
coarse aggregate was restricted to 10 mm in the mixture.

6.7.2. Compressive Strength


Figures (6.2) and (6.3) show the testing set-up and the EPS-LWC specimens used to
determine the compressive strength. Table (6.11) presents the recorded values of
compressive strength in EPS-LWC at seven different ages. The compressive strength of
different shape and size of the specimens, together with the effects of size and age are
discussed in the following sections.

The Avery (UK, Birmingham) hydraulic testing machine with capacity of 50 tonnes
was used to determine the compressive strength and modulus of elasticity of concrete.
According to the instructions of the testing machine, determination of the compressive
strength, the loading rates were different for the different shapes and sizes of the
specimens. For example, the cylindrical specimens with dimensions of 150 mm ¯300
mm were loaded in the testing machine under controlled loading at the rate of 0.3
MPa/sec until failure.

213
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

Table 6.11: Compressive strength of lightweight concrete mixture at different ages

Age (day) 3 7 14 21 28 56 91

Compressive strength (MPa) 14.49 25.38 27.93 29.62 30.96 32.33 33.3

Figure 6.2: Testing setup and different sizes and shapes of the specimens for compressive strength tests

The results show that the compressive strength at 3 and 7 days are 46.79% and 82.97%
respectively of the compressive strength at the age of 28 days. Also the corresponding
compressive strength values at ages 56 and 91 days are 4.4% and 7.64% higher than that
of 28 days respectively. In other words, the main development of compressive strength up
to 45% occurs at early age (3 days).

Equation (6.1) is proposed in this study to explain the relationship between the
compressive strength of EPS-LWC at different ages with respect to the strength at 28
days.


݂௖ᇱ ሺ‫ݐ‬ሻ ൌ ሺͲǤͳ͸͵͵ Ž ‫ ݐ‬൅ ͲǤͶͳ͹ሻ ݂௖ିଶ଼ Eq. (6. 1)


Where; ݂௖ᇱ ሺ‫ݐ‬ሻ is the compressive strength (MPa) at age t (days) and ݂௖ିଶ଼ is the
compressive strength (MPa) at age 28 days.

Figure (6.3) compares the predicted values against experimental values of the
compressive strength at various ages. Equation (6.1) underestimates the values for
compressive strength under the age of 28 days, while it slightly overestimates the
compressive strength for 56 and 91 days. However, the differences between the predicted
and experimental value are not significant.

214
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

Figure 6.3: Experimental vs. predicted values of compressive strength by Eq. (5.1)

6.7.2.1. Effects of shape and size of test specimens on the compressive strength

Different size and shapes of concrete specimens were used to determine the
compressive strength at different ages. Totally, 48 cylindrical specimens (18 cylinders
with 100mm¯200 mm, 18 cylinder specimens with 150mm¯300 mm and 12 cylindrical
specimens with 75¯150 mm dimensions) and 36 cubic specimens (18 cubes with 100
mm and 18 cubes with 150 mm dimensions) were used to investigate the effect of size
and shape of the specimen on the measured compressive strength of EPS-LWC. Figure
(6.4) shows the schematic presentation of the shape and size of the test specimens.

Figure 6.4: Different size and shapes of the compressive strength test specimens

The measured values of compressive strength at ages 3, 7, 14, 21, 28, 56 and 91 days
for three sizes of cylindrical specimens and 2 sizes of cubic specimens are presented in
Figure (6.5).

An overall comparison of the diagrams for the cubic and cylindrical specimens in
Figure (6.5) reveals that the compressive strength values measured by the cube specimens

215
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

are higher than those for cylindrical specimens. However, the difference in compressive
strength in the cylinder with 75mm¯150 mm dimension and the cube with 100 mm
dimension is not large. The cylinder with 150mm¯300 mm dimension has the lowest
compressive strength, while the cube with 100 mm dimension shows the highest
compressive strength in the majority of the test ages. By comparing the average values of
the compressive strength at all ages, the ratio of compressive strength of the 100 mm
cube specimen to the compressive strength of the 150 mm cube specimen is 1.056. The
ratio of the compressive strength in the 100 mm cube specimen to the compressive
strength of the 75mm¯150 mm, 100mm¯200 mm and 150mm¯300 mm cylindrical
specimens are 1.007, 1.11 and 1.16 respectively.

Figure 6.5: Compressive strength of LWC by different size and shape of test specimens

It is worth mentioning that the cylinder of 75mm¯150 mm is not recommended for


measuring of the compressive strength of concrete either in Australian or other
international codes of practice. However, considering the identical mix design, testing
methods and testing machines in all specimens, it can help evaluate the results by
comparing other sizes and shapes of EPS-LWC specimens.

6.7.3. Splitting Tensile Strength


Figure (6.6) shows the specimen and testing equipment used to measure the splitting
tensile strength of EPS-LWC in the laboratory conditions.

216
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

Figure 6.6: Specimen and test setup of splitting tensile strength

Table (6.12) presents the measured values of the splitting tensile strength at different
ages. Development of the splitting tensile strength with age is represented by the
proposed Equation (6.2) in this study.

Table 6.12: Splitting tensile strength of LWC mixture at different ages

Age (day) 3 7 14 21 28 56 91

Splitting tensile
1.44 2.3 2.6 2.78 2.89 2.98 3.08
strength (MPa)

݂௖௧ ሺ‫ݐ‬ሻ ൌ ሺͲǤͳͷ͸ Ž ‫ ݐ‬൅ ͲǤͶ͵Ͷͳሻ ݂௖௧ିଶ଼ Eq. (6. 2)

Where; ݂௖௧ ሺ‫ݐ‬ሻ is the splitting tensile strength (MPa) at age t (days), and ݂௖௧ିଶ଼ is the
splitting tensile strength (MPa) at age 28 days.

Figure (6.7) compares the predicted and experimental values of the splitting tensile
strength at various ages. Equation (6.2) gives underestimated values for the splitting
tensile strength under age 28 days, while it slightly overestimates the splitting tensile
strength for 56 and 91 days. However, the differences between the predicted and
experimental value are not significant.

217
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

Figure 6.7: Experimental vs. predicted values of splitting tensile strength by Eq. (5.2)

6.7.3.1. Effect of shape and size of test specimens on splitting tensile strength

A total of 36 cylindrical specimens of two different sizes (18 cylinders with


100mm¯200 mm and 18 cylinders with 150mm¯300 mm dimensions) were used to
measure the splitting tensile strength of EPS-LWC.

Figure (6.8) shows the tensile strength test results for cylindrical specimens with
100mm¯200 mm and 150mm¯300 mm dimensions. The cylinder with the smaller
diameter gives about 7% higher tensile strength. Despite using the same concrete mixture
the different sizes of cylinder specimens show different values of splitting tensile strength
at the corresponding ages. Development rate of the splitting tensile strength with age is
also influenced by the size of cylinder specimen.

Figure 6.8: Splitting tensile strength of cylindrical specimens


218
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

6.7.4. Modulus of Elasticity


Cylindrical specimens with dimension of 150mm¯300 mm are utilized to measure the
modulus of elasticity of LWC containing EPS beads at different ages. Three specimens
per age are tested in the concrete laboratory. Figure (6.9) shows the testing set-up and the
specimens used to measure the modulus of elasticity of EPS-LWC.

Figure 6.9: Specimen and test setup of modulus of elasticity

Table (6.13) shows the measured values of the modulus of elasticity of EPS-LWC at
different ages. The results show that the recorded values of the modulus of elasticity at 3,
7 and 91 days are 57%, 77% and 107% of the modulus of elasticity at age 28 days
respectively.

Table 6.13: Modulus of elasticity of LWC mixture at different ages

Age (day) 3 7 14 21 28 56 91

Modulus of elasticity (GPa) 11.63 15.68 17.65 19.48 20.32 21.1 21.89

Figure (6.10) shows development of the modulus of elasticity of LWC by age. This
development by age can be expressed by Equation (6.3). In fact, the equation represents
the age factor in which the ratio of the modulus of elasticity at different ages to the
modulus of elasticity at age 28 days is illustrated.

‫ܧ‬௖ ሺ‫ݐ‬ሻ ൌ ሺͲǤͳͶ͸͹ Ž ‫ ݐ‬൅ ͲǤͶ͸͸͸ሻ ‫ܧ‬௖ିଶ଼ Eq. (6. 3)

Where; ‫ܧ‬௖ ሺ‫ݐ‬ሻ is the modulus of elasticity (GPa) at age t (days) and ‫ܧ‬௖ିଶ଼ is the
modulus of elasticity (GPa) at age 28-days.

219
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

25

20

Modulus of elasticity (GPa)


15

10

0
0 20 40 60 80 100
Age (day)

Figure 6.10: Development of modulus of elasticity of lightweight concrete with age

Figure (6.11) compares the predicted values against the experimental values of the
modulus of elasticity at different ages. Equation (6.3) gives underestimated values for the
modulus of elasticity at age 28 days, while it overestimates the modulus of elasticity at 56
and 91 days. However, the predicted and experimental values are very close.

Figure 6.11: Experimental vs. predicted values of modulus of elasticity by Eq. (6.3)

220
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

6.7.5. Modulus of Rupture (Flexural Tensile Strength)


The modulus of rupture (flexural tensile strength) test of EPS-LWC was conducted by
a prism with dimensions of 100 mm × 100 mm × 350 mm under third-point loading. The
main objectives of this test were to determine the flexural toughness and first crack
strength of LWC. Figure (6.12) shows the specimen and testing equipment used to
measure the modulus of rupture (flexural tensile strength) of EPS-LWC.

Figure 6.12: Specimen and testing equipment of flexural tensile strength

Table (6.14) presents the measured modulus of rupture of EPS-LWC at different ages.
The ultimate rupture load, applied from the testing machine, is also given in Table (6.14).
According to the test results, the modulus of rupture at ages 3, 7 and 91 days are 69%,
81% and 105% of the modulus of rupture at age 28 days respectively. In other words,
EPS-LWC gains its flexural tensile strength mainly at the early ages.

Table 6.14: Modulus of rupture of lightweight concrete at different ages

Age (day) 3 7 14 21 28 56 91

Ultimate rupture load (KN) 9.94 11.66 13.52 14.06 14.35 14.58 15.08

Modulus of rupture (MPa) 2.982 3.498 4.056 4.218 4.305 4.374 4.524

Figure (6.13) shows the general development of the modulus of rupture of EPS-LWC
by age. This development by age can be expressed by the proposed Equation (6.4) in this
study. In fact, the equation represents the age factor in which the ratio of the modulus of
rupture at different ages in relation to the modulus of elasticity at the age of 28 days is
explained.

݂௥ ሺ‫ݐ‬ሻ ൌ ሺͲǤͳͲͶ͸ Ž ‫ ݐ‬൅ ͲǤ͸ʹሻ ݂௥ିଶ଼ Eq. (6. 4)

221
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

Where ݂௥ ሺ‫ݐ‬ሻ is the modulus of rupture (MPa) at age t (days) and ݂௥ିଶ଼ is the modulus
of rupture (MPa) at age 28-days.

Figure 6.13: Development of modulus of rupture of lightweight concrete by age

Figure (6.14) compares the predicted and experimental values of the modulus of
rupture at different ages. Equation (5.4) gives very close predicted values for the modulus
of rupture, especially over the age of 28 days.

Figure 6.14: Experimental vs. predicted values of modulus of rupture by Eq. (5.4)

To investigate the mid-span deflection of the test specimens under the third-point
loading, a transducer was attached to the prism specimen. The mid-span deflection was
222
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

the average of the ones detected by the transducers through the contact with brackets
attached to the prism specimen. Figure (6.15) compares the load-deflection diagram at
mid-span of the test prisms at ages 7, 28 and 56 days.

0.6

0.5

Deflection at midspan (mm)


0.4
7 days
0.3 28 days
56 days
0.2

0.1

0
0 5 10 15 20
Load (KN)

Figure 6.15: Load-deflection in mid-span of the flexural test specimens at different ages

The rupture load at 56 days is about 2% higher than that of 28 days; however, the mid-
span deflection at 28 days is about 25% higher. This difference might be explained as
result of the higher flexural stiffness of the specimen at 56 days. The rupture load and
maximum deflection at age 7 days are considerably below those of 28 days.

6.7.6. Drying Shrinkage


The drying shrinkage test was conducted on the EPS-LWC prisms with 75 mm × 75
mm × 280 mm dimensions subjected to air-drying in the laboratory controlled conditions.
The main objective of this test is to determine the shrinkage strain after a long period of
drying which will be used in the time-dependent deflection calculation of EPS-LWC
slabs. Also the interaction between the weight loss of specimens and the shrinkage strain
at different ages is investigated.

Figure (6.16) shows the specimen and test setup used to measure the shrinkage strain
of EPS-LWC in this study.

223
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

Figure 6.16: Shrinkage specimens and testing equipment

6.7.6.1. Curing and drying condition of shrinkage test at different ages

To study the drying shrinkage strain and weight loss of the shrinkage prisms, a total of
nine specimens in 3 series of A, B and C are investigated. To explore the effect of curing
duration on the shrinkage behaviour of specimens, each group containing 3 specimens
was subjected to specific curing conditions. Different conditions of initial curing, drying
and rewetting were applied to the specimens as illustrated in Figure (6.17). The relative
humidity and temperature of the environment were kept at 50%±5 and 23 ±2oC
respectively.

Figure 6.17: Initial curing, drying and rewetting conditions of shrinkage prisms

Table (6.15) shows the measured values of shrinkage strain and weight of the
specimens at different ages after one-week of initial curing. According to Table (6.15),
the shrinkage and weight measurement in the specimens of series A, B and C was
initiated at 7, 21 and 35 days after pouring the concrete in the moulds. The specimens of
series B were maintained in water between the days 35 and 49 after casting, so, their
shrinkage and weight were not measured in that period. The presented values in Table
(6.15) are the average of the 3 specimens in each series.

Figure (6.18) shows the development of shrinkage strain with time. Although the B
and C series have been stored for a longer time in the water, there is a significant
difference in the recorded strains, especially at early ages. It emphasizes the effect of the
224
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

longer curing time to reduce the drying shrinkage. Specimens of series C are subjected to
drying, followed by two weeks of being stored in water, with the shrinkage and weight
measurement beginning after the second stage of rewetting.

In the specimens of series B, which had a longer initial curing time than the other
specimens, there is a reduction in drying shrinkage (swelling) after the second stage of
rewetting. Swelling is defined as the expansion of concrete specimen due to water
absorption in the cavities. It can be concluded that the better initial curing improves the
capability of the specimen recovering the shrinkage strain, particularly at the early ages.

Table 6.15: Recorded shrinkage strain and weight of specimens at different ages

Time (day) after Shrinkage strain (με) Weight of specimen (gr)


initial curing Specimen series Specimen series
A B C A B C
0 0 0 0 3399 3354.9 3321
1 140 168 3367 3289
15 658 536 685 3319 3301 3240
18 711 605 744 3316 3296 3237
19 741 642 780 3314 3293 3235
20 771 673 805 3312 3291 3234
21 793 696 825 3311 3289 3231
29 858 887 3307 3228
33 904 942 3305 3226
42 963 299 1002 3298 3353 3219
47 978 452 1006 3297 3318 3218
49 1001 546 1031 3296 3313 3217
57 995 735 1042 3295 3302 3216
71 1010 858 1061 3294 3293 3213
78 1029 902 1081 3293 3290 3212
88 1058 993 1111 3291 3286 3210
103 1039 964 1094 3291 3283 3210
112 1053 984 1111 3292 3282 3210
132 1100 1043 1167 3291 3278 3208
159 1121 1092 1188 3289 3274 3206
201 1173 1158 1235 3286 3268 3202
300 1212 1192 1268 3271 3249 3188
393 1239 1212 1296 3256 3234 3173
425 1253 1215 1311 3248 3227 3166
450 1264 1216 1322 3245 3224 3162

225
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

According to Figure (6.18), the recorded shrinkage strain in all specimens is in the
range of 1200-1400 micro-strains. The average ultimate shrinkage strain in series B after
450 days is 96% of the average ultimate shrinkage strain of series A. The specimens of
series C represent relatively a higher shrinkage strain among all series.

Based on the commencement age of the measurement series C, the average ultimate
shrinkage strain after 450 days is 5% higher than that of series A specimens. The results
confirm that the major part of the shrinkage strain up to 50% occurs at early ages of the
concrete.

1400 1400

1200 1200

1000 1000

Shrinkage strain (με)


Shrinkage strain (με)

800 800

600 600

400 400
Series A Series B
200 200

0 0
0 50 100 150 200 250 300 350 400 450 0 50 100 150 200 250 300 350 400 450
Age (day) Age (day)

1400 1400

1200 1200

1000 1000
Shrinkage strain (με)

Shrinkage (με)

800 800

600 600

400 400 Series A


Series A Series B
200 200
Series C
0 0
0 50 100 150 200 250 300 350 400 450 0 50 100 150 200 250 300 350 400 450
Age (day) Age (day)

Figure 6.18: Shrinkage variation in series A and C and shrinkage and swelling in series B

Figure (6.19) shows changes in the weight of specimens at different ages. As a matter
of interest, the rewetting in specimen B has caused an increase in the total weight due to
the water absorption.

226
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

3450 3450

Series A Series B
3400 3400

3350 3350
Weight (gram)

Weight (gram)
3300 3300

3250 3250

3200 3200
0 50 100 150 200 250 300 350 400 450 0 50 100 150 200 250 300 350 400 450
Duration(day) Duration(day)

3450 3450
Series A
3400 Series C 3400
Series B
3350 3350 Series C
Weight (gram)

Weight (gram)
3300 3300

3250 3250

3200 3200

3150 3150

3100 3100
0 50 100 150 200 250 300 350 400 450 0 50 100 150 200 250 300 350 400 450
Duration(day) Duration(day)

Figure 6.19: Weight of shrinkage test prisms at different ages

Weight loss (ΔW) and the ratio to the initial weight (W0) of specimens (ΔW/W0) are two
principal factors in studying the effect of specimen weight on the shrinkage behaviour.
Considering the drying shrinkage, the main part of the weight loss of specimens is due to
the loss of water. On the other hand, water loss is strongly related to the relative humidity
of the environment which affects the shrinkage strain.

Figure (6.20) shows the weight loss of specimens at different ages. The weight loss
follows a similar trend in series A and C after about 100 days from casting. However, due
to the applied various conditions of drying and rewetting, the weight loss in series B is
completely different, especially after rewetting. Following a significant weight recovery
after rewetting, the rate of weight loss increases in specimen series B, but the final lost
weight is still less than that of specimen series A. The final weight loss in specimens of
series B and C are 50% and 43.5% below the weight loss of specimen A respectively.

227
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

200 200

Series A 160 Series B


160
120

Weight loss (gram)

Weight loss (gram)


120
80

80 40

0
40 0 50 100 150 200 250 300 350 400 450
-40
0
0 50 100 150 200 250 300 350 400 450 -80
Duration (day) Duration (day)

200 200
Series A
Series C 160 Series B
160
Series C
120
Weight loss (gram)

Weight loss (gram)


120
80

80 40

40 0
0 50 100 150 200 250 300 350 400 450
-40
0
0 50 100 150 200 250 300 350 400 450 -80
Duration (day) Duration (day)

Figure 6.20: Rate of weight loss of shrinkage specimens

Figure (6.21) shows the rate of weight loss of specimens at different ages. Despite
similar trends in specimens of series A and C, the rate of weight loss is entirely different
in specimen B. The rate of weight loss in all specimens tends to be zero after about 100
days. However, there is a sharp drop in the rate of weight loss in specimens A and B at
first two weeks. In the specimen B, which had a longer curing time, the difference
between the initial and ending rate of the weight loss is considerably less than that of
specimen A.

228
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

35 35
Series B
30 30

Rate of weight loss (gram/day)

Rate of weight loss (gram/day)


Series A 25 Series B
25
20
20
15
15
10
10
5
5 0
0 50 100 150 200 250 300 350 400 450
0 -5
0 50 100 150 Duration
200 250 (day) 300 350 400 450 Duration (day)

35 35

30 30 Series A
Rate of weight loss (gram/day)

Rate of weight loss (gram/day)


Series C 25 Series B
25
Series C
20
20
15
15
10
10
5
5 0
0 50 100 150 200 250 300 350 400 450
0 -5
0 50 100 150 Duration
200 250 (day) 300 350 400 450 Duration (day)

Figure 6.21: Variation of weight loss rate with age of specimens

Figure (6.22) shows the relationship between the shrinkage strain and weight loss of
the specimens. In specimen of series A, there is a large increment of shrinkage, increasing
the weight loss of specimens up to 120 grams. After that point, despite the increasing
weight loss, the shrinkage strain variation changes to a steady state. Specimen of series B
experiences a similar trend, however, the effect of the extended initial curing on the
weight recovery and the stability point of the shrinkage is evident in Figure (6.22).

Since the specimens had different initial weights before the drying shrinkage test, the
independent weight loss of specimens cannot be a good indicator of shrinkage behaviour.
Therefore, the ratio of weight loss (ΔW) of specimens to their initial weight (W0) is
investigated. Figure (6.23) shows the relationship between ΔW/W0 ratio and the shrinkage
strain of specimens at different ages.

229
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

1400 1400

1200 1200 Series B

Shrinkage and swelling strain (με)


1000
1000

Shrinkage strain (με)


800
800
600
600 400
Series A Series B
400 200
0
200
-80 -40 0 40 80 120 160
-200
0
0 40 80 120 160 -400
Weight loss (gram) Weight loss (gram)

1400 1400

1200
1200

1000
1000
Shrinkage strain (με)

Shrinkage strain (με)


800
800
600
Series A
600 400
Series C Series B
400 200 Series C
0
200 -80 -40 0 40 80 120 160
-200
0
-400
0 40 80 120 160
Weight loss (gram) Weight loss (gram)

Figure 6.22: Weight loss of specimens versus shrinkage strain

The general trend of diagrams is similar in Figures (6.22) and (6.23). However, the
effect of ΔW/W0 ratio on the shrinkage of series A is higher than other specimens.

1400 1400

1200 1200
1000
1000
800
Shrinkage strain (με)

Shrinkage strain (με)

800
600
600 Series A Series B
400
400 200

200 0
-2 -1 0 1 2 3 4 5
-200
0
0 1 2 3 4 5 -400
Weight loss / Initial weigth (%) Weight loss / Initial weigth (%)

1400 1400

1200 1200

1000
1000
800
Shrinkage strain (με)

Shrinkage strain (με)

800
600 Series A
600 Series C Series B
400
Series C
400 200

200 0
-2 -1 0 1 2 3 4 5
-200
0
0 1 2 3 4 5 -400
Weight loss / Initial weigth (%) Weight loss / Initial weigth (%)

Figure 6.23: Lost weight per initial weight versus shrinkage strain
230
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

Apart from the ultimate shrinkage strain value, the increasing rate of shrinkage strain
might have a significant effect on the performance of concrete structures, particularly at
the early ages of construction. In another hand, the rate of water loss (weight loss of
concrete member) can complicate the prediction of the shrinkage strain. Figure (6.24)
shows the relationship between the shrinkage rate and the rate of water loss in the test
specimens.

The trend in specimen A can be illustrated by a bilinear diagram, while the changes in
the shrinkage rate with different rates of weight loss are almost linear. The maximum
rates of weight loss and shrinkage rate in specimen B are about 10% and 25% of those for
specimen A respectively.

Since the significant part of the weight loss of specimen C happened in the first drying
period before the measurement, the weight loss and consequently, the rate of weight loss
are not comparable with those parameters in specimens A and B.

180 180
160 160
140 140
Shrinkage rate (με/day)

Shrinkage rate (με/day)

120 120
100 100
80 80
60 60

40 Series A 40 Series B
20
20
0
0
-10 -20 0 10 20 30 40
0 10 20 30 40
Rate of weight loss (gram/day) Rate of weight loss (gram/day)

180 180
160 160
140 140
Shrinkage rate (με/day)
Shrinkage rate (με/day)

120 120
100 100
80 80
60 60 Series A

40 Series C 40 Series B
20 Series C
20
0
0
-10 -20 0 10 20 30 40
0 10 20 30 40
Rate of weight loss (gram/day) Rate of weight loss (gram/day)

Figure 6.24: Rate of weight loss versus shrinkage rate

To study the effect of the increasing rate of the ΔW/W0 ratio (percentage of weight loss
per day) on the shrinkage rate, Figure (6.25) compares the relationship between these two
factors in all specimens. Similar to Figure (6.24), the relationship in specimens A and C
can be illustrated by a bilinear diagram and the behaviour of specimen B by a multi-linear
diagram. There is no particular trend in specimen C. The maximum rate of percentage of
231
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

weight loss in specimen B is about 10% and 25%, compared to specimens in series A.
This rate is in agreement with the ratio of the maximum rate of weight loss in Figure
(6.24).

180 180
160 160
140 140
Shrinkage rate (με/day)

Shrinkage rate (με/day)


120 120

100 100

80 80

Series A
60 Series B
60
40
40
20
20
0
0 -0.2 0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1 -20
Rate of ΔW/W0 (% / day) Rate of ΔW/W0 (% / day)

180 180
160 160
140 140
Shrinkage rate (με/day)

Shrinkage rate (με/day) 120


120
100 100

80 80
Series A
60 Series C
60 Series B
40 Series C
40
20
20
0
0 -0.2 0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1 -20
Rate of ΔW/W0 (% / day) Rate of ΔW/W0 (% / day)

Figure 6.25: Rate of lost weight per initial weight versus shrinkage rate

6.7.6.2. Proposed analytical models

The mathematical models are effective tools used to express the properties of materials
in absence of the accurate tests. These models are generally proposed on the basis of the
most-compatible of the previously developed models in the literature and are verified by
the experimental data. There are different models used in literature to predict the
shrinkage strain of concrete at different ages. Tables (4.2) and (4.3) in Chapter 4 show
some existing shrinkage models for LWC and conventional concrete.

Figure (6.26) compares the measured experimental data from the shrinkage specimens
of this study versus the predicted values of the existing models presented in Table (4.2).
There are obvious differences between the predictions of the models and measured
values. According to Figure (6.26), all investigated models underestimate the shrinkage
of EPS-LWC. The models in AS-3600(2009) [31] and Best and Polivka (1959) [32] give
more reasonable predictions, however, differences between the measured experimental
data and the prediction of AS-3600(2009) [31] is increasing by time. ACI-209R-92
232
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

(2008) [33] significantly underestimates the shrinkage strain of EPS-LWC at all ages,
while ACI-435-R95 (2003) [34] gives shrinkage values between the predictions of AS-
3600(2009) and ACI-209R-92 (2008).

1400 1400
Best and Polivka (1959)
AS-3600-09
1200 Predicted = Experimental 1200 Predicted=Experimental
Best fitting line with
Predcited shrinkage data (με)

Predcited shrinkage data (με)


Best fitting line with
1000 predicted data 1000 predicted data

800 800

600 600

400 400

200 200

0 0
0 200 400 600 800 1000 1200 1400 0 200 400 600 800 1000 1200 1400
Experimental shrinkage data (με) Experimental shrinkage data (με)
1400 1400
ACI-209-08 ACI-435-R95(2003)
1200 1200
Predicted= Experimental
Predcited=Experimental
Predcited shrinkage data (με)

Predcited shrinkage data (με)

1000 Best fitting line with 1000


predicted data Best fitting line with
predicted data
800 800

600 600

400 400

200 200

0 0
0 200 400 600 800 1000 1200 1400 0 200 400 600 800 1000 1200 1400
Experimental shrinkage data (με) Experimental shrinkage data (με)

Figure 6.26: Experimental data of EPS-LWC versus predicted shrinkage strain by different models

However, the recorded shrinkage strains in this study are well-predicted by the model
proposed by Best and Polivka (1959) [32]. Therefore, it is selected as the base of a new
empirical model for describing the development of shrinkage strain of EPS-LWC as
presented in Equation (6.5).

ߝ௦௛ ൌ ͳͺͺ ൈ ‫ ݐ‬଴Ǥଷଵହ Eq. (6. 5)

Where; ߝ௦௛ is the shrinkage strain (με) and t is the age of shrinkage specimen (days)
after one week moist curing.

233
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

Figure (6.27) compares the accuracy of the proposed model in this study with the
experimental data and predictions of the presented models in the literature as presented in
Table (4.2).

1400

1200

1000
shrinkage data (με

800

600

400

ACI-435-R95 AS-3600(2009)
200 ACI-209(2008) Best and Polivka (1959)
Experimental data Proposed model

0
0 50 100 150 200 250 300 350 400 450
Age after 7 days moist curing (day)

Figure 6.27: Experimental data of EPS-LWC versus predicted shrinkage strain by different models

6.7.6.3. Rate of weight loss with time in shrinkage specimens

According to Figure (6.21), there is a remarkable relationship between the rate of


weight loss and age of the specimen. This relationship can be expressed by the developed
Equation (6.6) in this study.

ܸοௐ ൌ ͶͳǤͻ͵͹ ൈ ‫ ݐ‬଴Ǥ଻ଽ଼ Eq. (6. 6)

Where; ܸοௐ is the rate of weight loss (gram/day) and t is the age of specimen (day)
after 7 days of moist curing. The correlation factor between the proposed model and the
measured data is ܴ ଶ ൌ ͲǤͻͻ͵. The correlation factors for the models proposed by Best
and Polivka, AS-3600(2009) and ACI-435-R95 are 0.79, 0.54 and 0.47 respectively.

6.7.7. Creep Coefficient of Lightweight Concrete


Standard concrete cylindrical specimens (150 mm ¯ 300 mm) were used to measure
the creep strain and coefficient of EPS-LWC in this study. The creep specimens and
LWC slabs were loaded simultaneously at the age of 14 days. There were two levels of
uniformly distributed sustained loading applied on the slabs. Magnitude of the sustained
234
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

loading at levels 1 and 2 were equal to 30% and 40% of the ultimate capacity of the slab
section at mid-span, respectively. Consequently, the creep specimens were loaded up to
30% and 40% of the ultimate compressive strength of the used EPS-LWC mixture. Each
creep rig consisted for 2 cylindrical specimens. Another companion cylindrical specimen
with the same dimensions was monitored to measure the shrinkage strain after 14 days
moist curing. The testing setup and measurement approach are presented in the codes
presented in Table (6.9). The environmental condition in the concrete laboratory was kept
at 23±2°C, and the Relative Humidity (RH) was kept with the tolerance of 50±5%.

Figure (6.28) shows the specimen and testing equipment used to measure the creep
coefficient of LWC containing EPS beads.

(a) (b)
Figure 6.28: Creep test rig at two different loading levels

Tables (6.16) and (6.17) present the measured strains in the cylindrical specimens up
to 225 days under the two different loading levels. Also, the creep coefficients at different
ages under two distinct levels of sustained loading are presented in the Tables (6.16) and
(6.17).

Figure (6.29) graphically shows changes of the creep coefficients for EPS-LWC
specimens under two levels of sustained loads.

The final creep coefficient after 225 days under loading levels 1 and 2 were 2.231 and
2.415 respectively. The maximum creep coefficient of specimens subjected to loading
level 2 is 8.2% higher than that of loading level 1, which is smaller than loading level 2.

235
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

Table 6.16: Creep strain and creep coefficient of lightweight concrete cylindrical specimen (load level 1)

Time after loading Creep strain (με) Creep coefficient


0
hr. 2 75 0.099
4 95 0.125
1 156 0.206
4 429 0.567
5 462 0.61
6 501 0.662
14 767 1.012
21 915 1.208
day 26 957 1.265
56 1238 1.635
87 1419 1.874
112 1495 1.975
140 1565 2.067
168 1611 2.128
196 1651 2.181
225 1689 2.231

Table 6.17: Creep strain and creep coefficient of lightweight concrete cylindrical specimen ( load level 2)

Time after loading Creep strain (με) Creep coefficient


0
hr. 2 115 0.119
4 155 0.161
1 245 0.254
4 609 0.631
5 668 0.692
6 729 0.754
14 1015 1.052
21 1180 1.222
day 26 1320 1.368
56 1665 1.725
87 1945 2.016
112 2053 2.127
140 2154 2.232
168 2225 2.306
196 2281 2.364
225 2330 2.415

The applied loading levels on the specimens were kept constant during the test. In
other words, the creep rigs were under constant sustained loading of lower and higher
levels for 225 days. The final creep coefficient in the creep rigs were in agreement with
the loading level and the ratio of final creep of rig under higher load level to that of lower
load level was compatible with the ratio of higher to lower loading magnitude. However,
variation of the creep coefficient with time was completely different from the constant
ratio of applied load. The ratio of creep coefficient in the rigs was increasing by time;
however, the creep coefficient under higher loading level experienced a relatively higher
rate of increment by time.

236
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

The general trend of creep coefficient of both creep rigs is descending by time. Despite
the constant ratio of applied loads during the test, the ratio of creep coefficients follows a
variable.

2.4
2.2
2
1.8
Creep coefficient 1.6
1.4
1.2
LOAD-30%
1
LOAD-40%
0.8
0.6
0.4
0.2
0
0 25 50 75 100 125 150 175 200 225 250
Age (day)

Figure 6.29: Creep coefficients of lightweight concrete under two different loading levels

25
Ratio of creep coefficients

20 Ratio of load levels

15
Ratio (%)

10

0
0 50 100 150 200 250
Age (day)

Figure 6.30: Comparing the ratio of creep coefficient at different ages

Figure (6.31) compares the changing ratio of creep strain at different ages in two creep
rigs with the constant ratio of applied loads on the specimens.

237
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

60

Ratio of creep strains


50
Ratio of load levels
40

Ratio (%)
30

20

10

0
0 50 100 150 200 250
Age (day)

Figure 6.31: Comparing the ratio of creep strain at different ages

According to figure (6.31), despite the constant ratio of applied loads, the ratio of creep
strain is following a descending rate from 53% during the first days to 37% at the age of
225 days. Comparing the decreasing ratio of creep strain and creep coefficient in Figures
(6.30) and (6.31), they follow a similar trend throughout the testing duration from the
loading date.

In addition, the final strain ratio is 27% higher than the ratio of applied loads, while,
the final creep coefficient ratio is close to the ratio of applied loads. In other words, the
ratio of the final creep coefficient after 225 days is more compatible with the ratio of the
applied load.

Together with the creep measurement, a companion cylindrical specimen of the same
dimensions was used to monitor the shrinkage behaviour. Table (6.18) shows the
shrinkage strain of EPS-LWC after 14 days of moist-curing measured by cylindrical
specimens at different ages.

Table 6.18: Shrinkage strain of cylindrical specimens (companion of creep specimens)

hr. Days
Age
2 4 1 4 5 6 14 21 26 56 87 112 140 168 196 225

Shrinka 4 5 14 25 29 31 48 60 67 103 115 122 129 134 137 140


ge (με) 0 7 3 9 3 7 7 7 1 2 7 4 7 1 7 5

Figure (6.32) compares the measured shrinkage strain with the measured creep strains
under the two different levels of sustained loading over a period of 225 days. The final
238
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

shrinkage strain is 87% and 60% of the final creep strain under the lower and higher
levels of loading, respectively.

2500

2000

1500
Strain (με)

1000 Creep-30%
Creep-40%

500 shrinkage

0
0 50 100 150 200 250
Age (day)

Figure 6.32: Development of creep and shrinkage strain of lightweight concrete by cylindrical
specimens

6.7.7.1. Proposed analytical models

The mathematical models are effective tools used to express the properties of materials
in absence of the accurate tests. These models are generally proposed on the basis of the
most-compatible of the previously developed models in the literature and are verified by
the experimental data. There are different models used in literature to predict the creep
coefficient of concrete at different ages. Table (4.5) in Chapter 4 shows some existing
creep models for lightweight concrete.

Equation (6.7) explains the relationship between the creep strain of EPS-LWC
cylindrical specimen and age under two different levels of sustained stress. The
coefficients of each loading level are indicated in Table (6.19).

ߝ௖௥ ൌ ͺͲͲߙ‫ݐ‬ఉ Eq. (6. 7)

Where; ߝ௖௥ is creep strain, α is the ratio of applied stress to the compressive strength of
concrete, β is the coefficient of time.

The age of specimen and magnitude of the sustained stress (fraction of the compressive
strength of concrete) are included in Equation (6.7). Considering the number of concrete

239
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

mixtures used in this study, inclusion of the other effective parameters of creep strain in
Equation (6.7) will not be practical. Figures (6.33) and (6.34) confirm the efficiency of
the proposed models in this study to predict the creep strain of EPS-LWC under different
levels of sustained stress.

Table 6.19: Coefficients of creep strain of EPS-LWC in Eq. (6.7)

Sustained stress (MPa) ߙ β

0.3݂௖ᇱ 0.3 00.375

0.4݂௖ᇱ 0.4 0.375

2500
Experimental vs predicted
Experimental = predicted
2000
Predicted creep strain (με)

Best fit line

1500

1000

500

0
0 500 1000 1500 2000 2500
Experimental creep strain (με)

Figure 6.33: Comparison of the experimental and predicted creep strain at load level 1

2500
Experimental vs predicted
Experimental = predicted
2000
Predicted creep strain (με)

Best fit line

1500

1000

500

0
0 500 1000 1500 2000 2500
Experimental creep strain (με)
Figure 6.34: Comparison of the experimental and predicted creep strain at load level 2

240
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

Equation (6.8) is proposed in this study to predict the creep coefficient of EPS-LWC
subjected to different levels of sustained stress. The coefficients of each loading level are
indicated in Table (6.20).

௧ೌ
‫׎‬ሺ‫ݐ‬ǡ ‫ݐ‬଴ ሻ ൌ ௕ା௧ ೎ ‫׎‬௨ Eq. (6. 8)

Where; ‫׎‬ሺ‫ݐ‬ǡ ‫ݐ‬଴ ሻ is creep coefficient, ‫׎‬௨ is ultimate creep coefficient, and a, b, c are
experimental parameters.

Table 6.20: Coefficients of creep strain of EPS-LWC in Eq. (6.7)

Sustained stress (MPa) a b c

0.3݂௖ᇱ 0.65 8 0.6

0.4݂௖ᇱ 0.75 10 0.72

Figures (6.35) and (6.36) confirm the efficiency of the proposed models in this study to
predict the creep coefficient of EPS-LWC under different levels of sustained stress.

2.5 Experimental vs predicted

Experimental = predicted
2
Predicted creep coefficient

Best fit line

1.5

0.5

0
0 0.5 1 1.5 2 2.5
Experimental creep coefficient
Figure 6.35: Comparison of the experimental and predicted creep coefficient at load level 1

241
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

2.5 Experimental vs predicted

Experimental = predicted
2

Predicted creep coefficient


Best fit line

1.5

0.5

0
0 0.5 1 1.5 2 2.5
Experimental creep coefficient

Figure 6.36: Comparison of the experimental and predicted creep coefficient at load level 2

6.7.7.2. Effect of shape and size of the test specimen on shrinkage strain

The shrinkage strain of EPS-LWC was measured by prism and cylindrical specimens
for 225 days. A prism with dimensions of 285mm¯75mm¯75 mm was moist-cured for
7 days and a cylindrical specimen with 150 mm diameter and 30 mm height was moist-
cured for 14 days. Figure (6.37) compares the shrinkage strain of prism and cylindrical
specimens up to the age of 225 days.

According to Figure (6.37), the ultimate shrinkage after 225 days in the cylindrical
specimen is 18.1% higher. Also, the rate of shrinkage strain in first two weeks is higher in
the cylinder. In other words, the shrinkage of the prism specimen is growing slowly
during the first two weeks, and then it experiences a sharp growth between the ages of 15
to 50 days. After 50 days, the shrinkage in the cylinder is increasing more rapidly with a
higher incremental rate until the age of 225 days. The difference between the records of
prisms and cylinders are increasing after 50 days remarkably.

The ratio of shrinkage at 225 days to 50 days is 1.19 in the prism and 1.40 in the
cylindrical specimen, respectively. It means a faster growing of shrinkage in the
cylindrical specimen comparing to the prism specimen.

242
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

1600

1400

1200

Shrinakge strain (με)


1000

800

600
Prism (7 days cured)
400
Cylinder (14 dasy cured)
200

0
0 25 50 75
100 125 150 175 200 225 250
Age (day)
Figure 6.37: Shrinkage strain of EPS-LWC by prism and cylindrical specimen

6.7.7.3. Volume/surface ratio

The ratio of specimen volume to the drying surface (V/S) is an important parameter in
studying the shrinkage behaviour of same mixture and similar curing conditions by
different shape and sizes of concrete specimens. The V/S ratio in prism and cylinder
specimens are 18.75 and 37.5 respectively. In other words, the hypothetical thickness in
the cylinder is twice that of the prism. The increased V/S ratio decreases the ultimate
shrinkage [35]. It is in good agreement with the recorded data up to the age 50 days.
However, it is in contrast with the final shrinkage value after 225 days.

6.7.8. Pull-Out Test


The pull-out test is performed to measure the bond stress between reinforcement and
concrete. The test specimens were a cylindrical specimen of 150mm ¯300mm concrete
with a steel bar of N12 in the middle. The specimens were cured in lime pre-saturated
water until the test day. The still-concrete bond stress was recorded at ages 3, 7, 14, 21,
28, 56 and 91 days. Three specimens were tested for each age.

Since, corrosion of steel bars in the water, especially the tests at 56 and 91 days
reduces the steel-concrete bond stress after casting, the steel bars and the concrete surface
surrounding the bar were cleaned by a wipe and covered by ant-burst oil-based spray.
Then all the specimens were kept in the water until the test age.

Figure (6.38) shows the general view of the universal pull-out test machine and the test
specimens with embedded steel bars.

243
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

Figure 6.38: Pull-out test specimens and testing equipment

6.7.8.1. Testing procedure

The loading range and rate of the testing machine must be selected properly according
to the expected failure load of the test specimen. The specimen is placed vertically on the
bearing plate provided with a central 2db cavity in the traction device, where the db is
steel bar diameter. The tension force (F) is applied to the lower extremity (the longer one)
while the apparatus for measuring the displacement ሺο଴ ሻ is placed at the upper extremity
of the bar.

6.7.8.2. Loading rate

According to the applied standard RILEM –RC6 (2004), the loading rate in Equations
(6.9) and (6.10) is determined for each bar diameter so that the rate of increments of the
bond stress is constant.

ி
ܸ௣ ൌ ௧ Eq. (6. 9)

ܸ௣ ൌ ͲǤͷ݀௕ଶ Eq. (6. 10)

Where; Vp is loading rate (N/sec), F is axial tension load (N), t is time (sec.) and db is
bar diameter (mm).

In addition to the included parameters in Equations (6.9) and (6.10), the length of the
steel bar inserted in the cylinder is another important factor in the pull-out test.

The inserted bar diameter at the centre of the cylinder was 12 mm (N12), hence, the
loading rate was selected as 72 N/sec or 4.32 KN/min. The embedded steel bar depth in
the cylinder specimens was 150 mm in the cylinder with total depth of 300 mm. The clear
cover all around the steel bar was 69 mm.

244
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

The specimen was loaded progressively up to the bond failure or the splitting of the
concrete cylinder, hence, the relation between tensile force (F) and the displacement (ο଴ ሻ
was determined.

The tension force F in the test were transformed by Equation (6.11) into bond stress ߬.
The bond stress has linear relationship with the ratio of the mean value of cylindrical

strength to the characteristic strength of concrete ( ௙೎೘ᇲ ).

ଵ ி ௙೎೘
߬ ൌ ହగ Ǥ ௗమ Ǥ ௙೎ᇲ
Eq. (6. 11)

Where; ݂௖௠ ൌ ͵Ͳ‫ʹݎ݋‬ͷǤͷis the mean value of the cylindrical specimen (N/mm2) and
݂௖ᇱ = characteristics compressive strength of test specimens (N/mm2).

As presented in Equation (6.12), in the pull-out test, the bond stress is also dependent
on the slip of reinforcement due to the tensile force F.

߬ ൌ ݂ሺ‫ݏ‬ሻEq. (6. 12)

Where; s is the slip measured under the tension force F.

The mean curve of all the individual test results is thus obtained for assessing of the
bond stress. Details of the specimen dimensions and the obtained test results are given in
Table (6.20)

Figure (6.39) shows the general trend of load increment with time in the pull-out test
specimens at different ages. The loading rate in all specimens is constant and the load-
time diagram is linear up to the failure point in the specimens of all ages. The failure
loading values increase by increasing the age of the specimen.

245
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

60
7 days
50 28 days
56 days
40

Load (kN)
30

20

10

0
0 200 400 600 800
Time (sec)
Figure 6.39: Failure load of specimens in pull-out test at different ages (loading rate=72N/sec)

Table 6.21: Specimen dimensions and variables of pull-out test

Age Hcyl dcyl Fu Slip ld db ࢌᇱࢉ Presence of


Cylinder
(day) (mm) (mm) (KN) (mm) (mm) (mm) (MPa) cracking
1 300 150 32.27 32 150 12 Un-cracked
3 2 300 150 34.81 35 150 12 14.49 Un-cracked
3 300 150 33.48 33 150 12 Un-cracked
1 300 150 43.94 26 150 12 Un-cracked
7 2 300 150 35.06 25 150 12 25.38 Un-cracked
3 300 150 41.52 25 150 12 Un-cracked
1 300 150 44.56 23 150 12 Un-cracked
14 2 300 150 43.25 20 150 12 27.93 Un-cracked
3 300 150 43.25 21 150 12 Un-cracked
1 300 150 45.15 20 150 12 Un-cracked
21 2 300 150 47.25 14 150 12 29.62 Un-cracked
3 300 150 44.32 19 150 12 Un-cracked
1 300 150 46.521 19 150 12 Un-cracked
28 2 300 150 50.199 8 150 12 30.96 Cracked
3 300 150 46.241 18 150 12 Un-cracked
1 300 150 44.67 17 150 12 Cracked
56 2 300 150 56.925 8 150 12 32.33 Cracked
3 300 150 53.525 9 150 12 Cracked
1 300 150 54.14 14 150 12 Cracked
91 2 300 150 53.54 13 150 12 33.3 Cracked
3 300 150 53.26 10 150 12 Cracked

246
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

In Table (6.20), ld and db are the embedded length and diameter of the tension bar in
middle of the cylindrical specimen.

6.7.8.3. Load-displacement behaviour of the test specimens

Figure (6.40 compares the load-displacement behaviour of pull-out test specimens at


ages 28 and 56 days. The failure mode in 56 days is brittle, while the specimen at age 28
days shows a ductile failure. In the 28 days specimen, the steel bar is pulling out without
any cracking in the concrete cylinder, while in the specimen tested at 56 days, slipping of
the steel bar occurs with splitting cracking of the cylinder.

60
28 days
50 56 days

40
Load (kN)

30

20

10

0
0 20 40 60 80
Displacement (mm)

Figure 6.40: Brittle and ductile failure of pull-out test specimens at 28 and 56 days

6.7.8.4. Effect of loading rate on the failure load

To investigate the effect of loading rate on the ultimate pull-out load, two series of
specimens at the age of 7 days were subjected to the different loading rates. First loading
was very slow at the rate of 2.5 N/sec while second loading rate was standard rate of 72
N/sec according to RILEM–RC6 (2204). As shown in Figure (6.41) low rate application of
load on the specimen increases the bond stresses between the steel bars and concrete.
Also, gradual slipping of the bars changes the rate of applied load especially, at the first
minutes of the loading.

The ratio of applied loads is 28.8 (= 72/2.5) however, the ultimate load at the low rate
is about 10% higher than the maximum tensile load at a standard rate.

247
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

50
loading rate=72 N/sec
45 Loading rate=2.5 N/sec
40
35
30

Load (kN)
25
20
15
10
5
0
0 500 1000 1500 2000 2500 3000
Time (sec)
Figure 6.41: Comparison of failure load by different loading rates

6.7.8.5. Bond-slip relationship

Table (6.21) shows the final bond between the steel and concrete and the measured
final slip values at different ages.

Table 6.22: Bond stress and corresponding slip values of test specimens at different ages

Age (days) 3 7 14 21 28 56 91

Bond stress ߬ (N/mm2) 14.83 17.77 19.33 20.16 21.08 22.87 23.73

Slip s (mm) 33.3 25.3 21.3 17.7 15 11.3 12.3

As presented in Tables (4.10) to (4.12) in Chapter 4 of this study, there is a power type
relationship between the bond stress and the bar slip. Accordingly, Equation (6.13) is
proposed in this study to describe the relationship between the bond-stress and the slip in
a reinforced lightweight concrete member under direct tension.

߬ ൌ ͸ʹǤͻʹͺ‫ି ݏ‬଴Ǥଷଽଽ Eq. (6. 13)

Where; “τ” is bond stress (MPa) and “s” is the slip of steel bar under tension (mm).

248
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

6.7.9. Compressive Stress-Strain Curve of Lightweight Concrete Containing


Expanded Polystyrene Beads
To obtain the stress-strain curves, the cylindrical specimens were made of EPS-LWC
were subjected to compression tests with a carefully controlled displacement rate. For
each age three cylindrical specimens were tested. The final stress-strain curve for each
age was taken as the average of three curves of tree cylindrical specimens. The
compressive stress-strain curves at the ages of 14 and 28 days are presented in Figure
(6.42).

35
14 days
30
28 days
Compressive stress (Mpa)

25

20

15

10

0
0 0.001 0.002 0.003 0.004 0.005
Strain
Figure 6.42: Stress-strain diagram of the LWC mixture at different ages

According to Figure (6.42), the general trend of compressive stress-strain variation in


the EPS-LWC is similar at two different ages. However, due to a lower elasticity of
hardened concrete at 14 days, the specimen shows higher strain values under the
compressive stress compared to the specimen at 28 days. The ascending part of the
Compressive Stress-Strain Curve (CSSC) at both ages are comparable, however, the
descending part is different. Additionally, although the peak stress point at 14 days is
lower than that of 28 days, however, the ultimate strain at both ages are tending to a
similar value.

6.7.10. Energy Absorption under Compression


To calculate the absorbed energy per unit volume of the cylindrical specimens under
compression, the Equation (6.14) is utilized. In fact, the absorbed energy is the area under
the compressive stress-strain curve up to the failure load.

249
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads


‫ܩ‬௖ ൌ ‫׬‬଴ ೠ ߪǤ ݀ߝEq. (6. 14)

According to Figure (6.42), the ultimate strain in calculation of the stored energy at
different ages was taken as 0.0042. Table (5.21) shows Gc values at different ages. Figure
(6.43) also compares the absorbed energy in compressive cylinders at different ages. Both
Table (6.22) and Figure (6.43) clearly show the increment of Gc by increasing the age of
concrete specimens.

Table 6.23: The energy absorbed in cylindrical specimens under compression

Age (days)

3 7 14 21 28 56 91

Gc (MPa) 0.02185 0.03892 0.04735 0.05477 0.0632 0.06856 0.07732

0.09

0.08

0.07
Dissipated energy (MPa)

0.06

0.05

0.04

0.03

0.02

0.01

0
0 20 40 60 80 100
Age (day)

Figure 6.43: The energy stored under compression at different ages

This development of absorbed energy by age can be expressed by Equation (6.15). In


fact, the equation represents the age factor at which the ratio of the modulus of elasticity
at different ages to the modulus of elasticity at the age of 28 days is illustrated.

‫ܩ‬௖ ሺ‫ݐ‬ሻ ൌ ሺͲǤͲͳ͸ ݈݊ ‫ ݐ‬൅ ͲǤͲͲ͸ͳሻ ‫ܩ‬௖ିଶ଼ Eq. (6. 15)

250
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

Where; ‫ܩ‬௖ ሺ‫ݐ‬ሻ is the stored energy per unit volume of the specimen (MPa) at the age of
t (days) and ‫ܩ‬௖ିଶ଼ is stored energy per unit volume of the specimen (MPa) at the age of
28 days.

Figure (6.44) shows the predicted versus experimental values of the energy absorptions
at different ages. According to Figure (6.40), the proposed time-dependent relationship is
in good agreement with the experimental results.

0.1
Experimental data

Predicted=Experimental
0.08
Predicted energy absorption (MPa)

Best fitting line with


predicted data

0.06

0.04

0.02

0
0 0.02 0.04 0.06 0.08 0.1
Experimental enery absorption (MPa)

Figure 6.44: Predicted versus experimental energy absorption under compression

Variation of stored energy per unit volume under compression with the strain changes
at different ages is shown in Figure (6.45). The Gc-ε diagrams in Figure (6.45) indicate
the higher rate of Gc increment with strain by increasing the age of the specimen; i.e.
older specimens demonstrate a quicker development of stored energy in the unit volume
of the specimen.

251
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

0.09
3 days
0.08
7 days

Stored energy per unitt volume (MPa)


0.07 14 days
0.06 21 days
28 days
0.05
56 days
0.04 91 days
0.03

0.02

0.01

0
0 0.001 0.002 0.003 0.004 0.005
Strain (mm/mm)

Figure 6.45: Stored energy under compression versus strain of EPS-LWC at different ages

Figure (6.46) shows the stress-strain relationship in the elastic range of the ascending
part of the CSSC at different ages of EPS-LWC specimens. By increasing the
compressive strength of the concrete, the corresponding elastic ranges develop with
different slopes. In other words, the increased compressive strength of concrete with age
increases the capacity of energy absorption (the area under the stress-strain curve) of the
specimens. The slope of diagrams increases rapidly at higher ages, especially after 14
days. Consequently, the rate of the stored energy increment after 14 days are higher than
the early ages of concrete.

14
3de
12 7de
14de
10 21de
28de
Stress (σ), MPa

8 56de
91de
6

0
0 0.0001 0.0002 0.0003 0.0004 0.0005 0.0006 0.0007
Strain (ε) , mm/mm

Figure 6.46: Compressive stress-strain variation of the LWC mixture in elastic range at different ages

252
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

6.7.11. Mechanical Properties of Steel Bars


The experimental part of this study is to investigate the time-dependent deflection of
reinforced concrete slabs. Details of the experimental program will be described in the
next chapters. To reinforce the concrete section, 4N12 steel bars have been embedded in
each slab. To enhance the accuracy of the study, the steel bars have been tested, and exact
values of mechanical properties of the steel bars have been determined. Mechanical
properties of the steel bar, including the yield stress, modulus of elasticity and poison
ration are given in Table (6.23). Figure (6.47) also presents the tensile stress-strain curve
of the N12 steel bars.

Table 6.24: Mechanical properties of N12 steel bar

Property Yielding stress, fy (MPa) Modulus of elasticity Es Poisson ratio (ν)


(kN/m2)
Measured value 538 200553150 0.25

700

600
Tensile stress(MPa)

500

400

300

200

100

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
Strain

Figure 6.47: Tensile Stress- strain curve of the N12 steel bars

6.7.12. Relationships between Compressive Strength and Mechanical Properties


The compressive strength can illustrate mechanical properties of hardened concrete.
There is a wide range of empirical relationships and models in literature to predict
different mechanical properties from compressive strength. Chapter Three of this study
presented some existing models for mechanical properties of LWC, CC, and SCC.

Table (6.24) summarizes the mechanical properties of EPS-LWC. Equations (6.14) to


(6.15) are proposed based on regression analysis of the experimental data to estimate the
splitting tensile strength, modulus of elasticity and modulus of rupture of EPS-LWC

253
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

based on compressive strength. The models presented in Chapter Three have been
selected as the basis of these equations.

Figures (6.48) to (6.51) compare the predicted versus experimental values of splitting
tensile strength, modulus of elasticity and modulus of rupture respectively.

Table 6.25: Mechanical properties of EPS-LWC at different ages

Age (day) 3 7 14 21 28 56 91

Compressive strength (MPa) 14.49 25.38 27.93 29.62 30.96 32.33 33.3

Splitting tensile strength (MPa) 1.44 2.3 2.6 2.78 2.89 2.98 3.08

Modulus of Elasticity (GPa) 11.63 15.68 17.65 19.48 20.32 21.1 21.89

Modulus of rupture (MPa) 2.982 3.498 4.056 4.218 4.305 4.374 4.524

6.7.12.1. Splitting tensile strength

Equation (6.16) is proposed to predict the splitting tensile strength of EPS-LWC at


different ages. The obtained correlation factor (R2) shows a good agreement between the
measured values and predicted values of the splitting tensile strength by the proposed
model.

଴Ǥଽଵ଻଺
݂௧௖ ൌ ͲǤͳʹʹ͹ ൈ ݂௖ᇱ ܴ ଶ ൌ ͲǤͻͻ͸ Eq. (6. 16)

4
Experimental data

Predicted=Experimental

3 Best fitting line with


Predicted ftc (MPa)

predicted data

0
0 1 2 3 4
Experimental ftc (MPa)
Figure 6.48: Predicted versus experimental splitting tensile strength of EPS-LWC

254
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

Together with the high correlation factor, according to Figure (6.48), the proposed
model estimates the splitting tensile strength from compressive strength with high
accuracy.

6.7.12.2. Modulus of elasticity

Equation (6.17) is proposed to predict the modulus of elasticity of EPS-LWC at


different ages.

଴Ǥ଻ହଶଷ
‫ܧ‬௖ ൌ ͳǤͷͲͶͷ ൈ ݂௖ᇱ ܴ ଶ ൌ ͲǤͻͺ͹Eq. (6. 17)

Along with the high correlation factor, according to Figure (6.49), the proposed model
of predicting of the modulus of elasticity from the compressive strength is in good
agreement with the experimental results.

25
Experimental data

Predicted=Experimental
20
Best fitting line with
Predcited Ec (GPa)

predicted data
15

10

0
0 5 10 15 20 25
Experimental Ec (GPa)
Figure 6.49: Predicted versus experimental modulus of elasticity of EPS-LWC

6.7.12.3. Modulus of rupture

Equation (6.18) is proposed to predict the modulus of rupture (flexural tensile strength)
of EPS-LWC at different ages.

଴Ǥହ଴଴ଷ
݂௥ ൌ ͲǤ͹͸ʹͷ ൈ ݂௖ᇱ ܴ ଶ ൌ ͲǤͻͳͻ͸ Eq. (6. 18)

255
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

According to Figure (6.50), the proposed model of predicting of the modulus of


rupture from the compressive strength is in good agreement with the experimental results.
Additionally, the high correlation factor clearly indicates efficiency of the proposed
model.

5
Experimental data

Predcited=Experimental
4
Predcited fr (MPa) Best fitting line with
predcited data
3

0
0 1 2 3 4 5
Experimental fr (MPa)

Figure 6.50: Predicted versus experimental modulus of rupture of EPS-LWC

The correlation factor (R2) between the predicted and experimental values of splitting
tensile strength, modulus of elasticity and modulus of rupture are presented in Table
(6.25). According to R2 values, the proposed models in Equation (6.14) to (6.16) for
prediction of the splitting tensile strength, modulus of elasticity and modulus of rupture
are precise and in good agreement with the experimental results.

Table 6.26: Correlation factor between the predicted and experimental mechanical properties

Property Equation R2
Splitting tensile strength 5.7 0.99
Modulus of elasticity 5.8 0.95
Modulus of rupture 5.9 0.92

6.7.12.4. Ultimate pull-out load with respect to the compressive strength

Figure (6.51) shows the relationship between the compressive strength of the LWC
mixture and the ultimate pull-out load on the cylindrical specimens. Equation (6.19) is
developed to explain the relationship between the ultimate pull-out load and the
compressive strength of EPS-LWC.

256
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

‫ ܨ‬ൌ ͲǤͻͻʹͷ݂௖ᇱ ൅ ͳ͹ǤͶͺͳܴ ଶ ൌ ͲǤͻ͸Eq. (6. 19)

Where; F is ultimate pull-out load (KN) and ݂௖ᇱ is compressive strength (MPa).

60

55

Ultimate pull-out load (KN)


50

45

40

35

30

25

20
10 15 20 25 30 35
Compressive strength (MPa)

Figure 6.51: Ascending ultimate pull-out load by age development of compressive

6.7.12.5. Compressive stress-strain relationship

The existing models to estimate the stress-strain diagram of different types of concrete
under compression are broadly discussed in Chapter Three of this study. Based on the
existing models presented in Chapter Three, a new model in Equation (6.20) is developed
to predict the compressive stress-strain curve of EPS-LWC. In the proposed model, the
Equations (6.16) to (6.18) proposed for compressive strength and modulus of elasticity,
are also utilized.

୤ᇲౙ ǤஒሺகȀகబ ሻ
ߪ௖ ൌ Ǣ Ͳ ൑ ɂ ൑ ɂ୫ୟ୶ Eq. (6. 20)
ൣሺஒିଵାሺகȀகబ ሻಊ ൧

ͳ
Ⱦൌ 
݂ᇱ
ͳ െ ௖ 
‫ܧ‬଴ ߝ଴

ˆୡᇱ Ǥ ȾሺɂȀɂ଴ ሻ
ߪ௖ ൌ ݂݅ɂ ൐  ɂ଴ 
ሾሺȾ െ ͳ ൅ ሺɂȀɂ଴ ሻ୬ஒ ሿ
݊ ൌ ͲǤͲͲͳʹܲଶ െ ͲǤͲͺͶ͸ܲ ൅ ͳǤͷͷͳ

‫ܿ݁ݏܧ‬ ݂௖ᇱ
ܲൌ ‫ ܿ݁ݏܧ‬ൌ 
‫ܧ‬଴ ɂ଴
257
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

ɂ଴ ൌ ͸Ǥͷʹ݂௖ᇱ ൈ ͳͲିହ ൅ ͳǤ͸͵Ͷ ൈ ͳͲିସ 

଴Ǥ଻ହଶଷ
‫ܧ‬଴ ൌ ͳǤͷͲͶͷ ൈ ݂௖ᇱ 

Where;

ɐୡ ǡ ɂ : Stress and strain of concrete

ˆୡᇱ ǡ ɂ଴ : Maximum stress and corresponding strain of concrete

β: Material parameter

‫ܧ‬଴ : Initial tangent modulus of elasticity

Figure (6.53) compares the measured v stress-strain under the two different levels of
compressive stress and the predicted values of creep strain by the proposed model in the
Equation (6.20).

0.005

0.004
Predicted strain (με)

0.003

0.002

0.001

0
0 0.001 0.002 0.003 0.004 0.005
Experimental strain (με)
Figure 6.52: Comparison of the measured and predicted stress-strain variation under compression

258
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

6.8. COMPARING THE MECHANICAL PROPERTIES OF


LIGHTWEIGHT, SELF-COMPACTING AND CONVENTIONAL
CONCRETE

6.8.1. Mixture Design


The experimental results of this study are compared with the previously conducted
experiments on self-compacting concrete slabs by Aslani (2014) [1] and conventional
concrete slabs by Nejadi (2005) [2].

The mixture design and components are different in these types of concrete. In total,
one mixture for LWC slabs and one mixture for CC slabs were used. For eight SCC slabs,
two normal SCC mixtures (N-SCC), two steel fibre (D-SCC), two polypropylene fibre
(S-SCC) and two hybrid fibre (DS-SCC) mixtures have been considered in the test
program.

Table (6.27) explains the mixture proportions and density of LWC and SCC mixtures.
Type, volume and aspect ratio of the fibres that influence the mechanical properties of
SCC mixtures are also illustrated in Tables (6.28). The mixture design of CC with density
of 2450 kg/m3 is not reported in the reference.

259
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

Table 6.27: Mixture design of self-compacting concrete [1] and lightweight concrete

Constituents N-SCC D-SCC S-SCC DS-SCC LWC CC

Cement (kg/m3) 160 160 160 160 250

Fly Ash (kg/m3) 130 130 130 130

GGBFS * (kg/m3) 110 110 110 110

Cementitious content (kg/m3) 400 400 400 400

Water (kg) 208 208 208 208 150 lit/m3

Water Cementitious ratio 0.52 0.52 0.52 0.52 0.6

Aggregate (kg/m3)

Coarse sand 660 660 660 660 620

fine sand 221 221 221 221

Coarse aggregate 820 820 820 820 800

Light Aggregate (BST) (lit/m3) 300

Admixtures (lit/m3)

Super plasticizer 4 4.86 4.73 4.5

Viscosity modifying agent 1.3 1.3 1.3 1.3

High range water reducing 1.6 1.6 1.6 1.6 2


agent
Fibre content (kg/m3)

Steel - 30 - 15

Polypropylene - - 5 3

Density 2340 2274 2330 2385 2000 2450

*Ground Granulated Blast Furnace Slag (GGBFS)

Table 6.28: Physical and mechanical properties of fibres in self-compacting concrete mixture [1]

Aspect Tensile Elasticity


Trade Density Length Diameter
Type Ratio strength modulus
name (kg/m3) (l) (db)
(l/d) (MPa) (GPa)
Drami
Steel x RC-80/60- 7850 60 0.75 80 1050 200
BN
Polypropylene Synmi x 65 905 65 0.85 76.5 250 3

260
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

6.8.2. Early-Age Mechanical Properties


In general, the existing models and empirical relationships in the literature provide
relatively inaccurate values of the mechanical properties of concrete. Besides, the models
used to predict the early age characteristics of concrete especially SCC and LWC are very
rare in the literature. Therefore, compressive strength and modulus of elasticity of the
mixtures were measured by the companion test specimens in the laboratory conditions at
two ages of 14 and 28 days for all different types of concrete. Table (6.29) shows the
measured compressive strength and modulus of elasticity of EPS-LWC, SCC and CC
mixtures at 14 and 28 days.

According to test results, all mixtures are in the range of normal strength concrete.

Table 6.29: Mechanical properties of self-compacting concrete [1], lightweight concrete and
conventional concrete [2] mixtures

Property Age (day) N-SCC D-SCC S-SCC DS-SCC LWC CC

14 29.05 34.3 32.45 38.1 27.93 18.3


݂௖ᇱ ሺ‫ܽܲܯ‬ሻ
28 33.3 38 38.1 45 30.96 24.8

14 32.24 29.14 29.68 31.26 17.65 22.82


‫ܧ‬௖ ሺ‫ܽܲܩ‬ሻ
28 35.39 35.76 35.76 36.1 20.32 24.95

Figures (6.53) and (6.54) compare the measured values of compressive strength and
modulus of elasticity of the mixtures at ages 14 and 28 days respectively. Development
rate of the compressive strength and modulus of elasticity with time in the mixtures are
also compared in Figures (6.55) and (6.56) respectively.

261
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

50
14 days 28 days
45

Compressive strength (MPa)


40

35

30

25

20

15
N-SCC D-SCC S-SCC DS-SCC LWC NC
Mixtures

Figure 6.53: Compressive strength of EPS-LWC, SCC and CC mixtures at 14 and 28 days

40
14 days 28 days
Modulus of Elasticity (GPa)

35

30

25

20

15
N-SCC D-SCC S-SCC DS-SCC LWC NC
Mixtures

Figure 6.54: Modulus of elasticity of EPS-LWC, SCC and CC mixtures at 14 and 28 days

N-SCC S-SCC D-SCC


DS-SCC LWC NC
50
Compressive strength (MPa)

45
40
35
30
25
20
15
10 15 20 25 30
Test age

Figure 6.55: Increasing rate of compressive strength in EPS-LWC, SCC and CC mixtures

262
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

N-SCC D-SCC S-SCC


DS-SCC LWC NC
40

Modulus of elasticity (GPa)


35

30

25

20

15
10 15 20 25 30
Test ages

Figure 6.56: Increasing rate of modulus of elasticity in EPS-LWC, SCC and CC mixtures

In Figures (6.55) and (6.56), the presented lines between two points (14 and 28 days)
are part of general development diagrams of the compressive strength and modulus of
elasticity with time. The general relationship between the age and mechanical properties
of concrete is not necessarily linear.

According to Figures (6.53) to (6.56), the average increasing ratio of the modulus of
elasticity in CC, SCC and EPS-LWC are 1.10, 1.17 and 1.10 respectively. The
development ratio of the compressive strength in CC, SCC and EPS-LWC is 1.35, 1.15
and 1.10 respectively. Overall, the early-age properties and increasing rate in SCC
mixtures, particularly in presence of the fibres are greater than the corresponding
properties of CC and LWC mixtures.

The compressive strength and modulus of elasticity are the main mechanical properties
of concrete affecting the deflection behaviour of reinforced concrete slabs. However,
other mechanical properties can influence the deflection behaviour of the slabs also. The
tensile strength of concrete under flexural loading is an effective parameter in the
cracking and deflection behaviour of EPS-LWC, SCC and CC slabs. Figure (6.57)
compares the modulus of rupture (flexural tensile strength) of CC, LWC and SCC
mixtures at different ages.

263
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

Modulus of rupture (MPa)


6

3
N-SCC D-SCC
2
S-SCC DS-SCC
1 LWC CC

0
0 20 40 60 80 100
Age (day)

Figure 6.57: Comparison of the modulus of rupture in LWC, CC and SCC mixtures

According to Figure (6.58), EPS-LWC poses the lowest values of the modulus of
rupture, especially after age 28 days. However, SCC mixture with and without fibre
reinforcing are in the top range of modulus of rupture in almost all ages.

6.8.3. Time-Dependent Properties of Conventional, Self-Compacting and


Lightweight Concrete
6.8.3.1. Shrinkage strain

Creep and shrinkage are the two main parameters which influence the time-dependent
cracking and deflection behaviour of reinforced concrete slabs. The shrinkage strain is
monitored for about 500, 250 and 400 days in EPS-LWC, SCC and CC specimens
respectively. Additionally, the creep strain of EPS-LWC, SCC and CC specimens were
measured for about 250 days under two distinct levels of compressive stress.

Figure (6.58) compares the shrinkage strain of EPS-LWC, SCC and CC mixtures up to
240 days. Table (6.30) summarizes the final free shrinkage strain of CC, LWC and CC
specimens after 240 days from casting date.

Table 6.30: Free shrinkage strain of CC, LWC and SCC mixtures

CC LWC N-SCC D-SCC S-SCC DS-SCC

Shrinkage strain (με) 784 1180 870 844 823 882

264
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

1200

1000

Shrinkage strain (με) 800

600
LWC
400 CC
N-SCC
D-SCC
200
DS-SCC
S-SCC
0
0 50 100 150 200 250
Age (days)
Figure 6.58: Comparison of shrinkage strain in EPS-LWC, SCC and CC mixtures

According to Figure (6.58) and Table (6.30), the shrinkage strain in LWC is
considerably higher than that of the CC and SCC mixtures. The free shrinkage strain in
LWC and N-SCC mixtures is 51% and 10% greater than that of the CC mixture. The
porous (permeable) nature of the lightweight concrete, relatively higher ratio of water to
cement and high rate of the water loss cause higher values of the drying shrinkage in
lightweight concrete.

6.8.3.2. Creep strain

Figure (6.59) compares the creep coefficient in EPS-LWC, SCC and CC specimens
under sustained compressive stress. The creep strain is measured by standard cylindrical
specimens in all concrete types. The applied stress is equal to 40% of the compressive
strength of the companion cylinders in each type of concrete. The creep strain under
compressive stress equal to 30% of the compressive strength of cylinders is not reported
for SCC and CC mixtures. However, the creep strain and creep coefficient of EPS-LWC
under 30% and 40% of the compressive strength is comprehensively explained in section
(6.7.8) of this chapter.

265
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

2.5

Creep coefficient
1.5

CC-40%
1
LWC-40%
N-SCC-40%

0.5 D-SCC-40%
S-SCC-40%
DS-SCC-40%
0
0 50 100 150 200 250
Age (day)
Figure 6.59: Creep coefficient of CC, SCC and EPS-LWC specimens loaded at age 14 days

According to Figure (6.59) the creep coefficient in EPS-LWC is considerably higher


than that of SCC and CC specimens under the same level of sustained compressive stress.

Table (6.31) summarizes the creep coefficient of EPS-LWC, SCC and CC mixtures at
240 days after casting.

Table 6.31: Creep coefficient of CC, LWC and SCC mixtures after 240 days sustained loading

CC LWC N-SCC D-SCC S-SCC DS-SCC

Creep coefficient 1.5 2.415 1.96 1.86 1.82 1.45

According to Figure (6.59), the creep coefficient in EPS-LWC and CC reach the
highest and lowest recorded values after 240 days. The creep coefficient at 240 days in
LWC and N-SCC mixtures is 1.61 and 1.3 times higher than that of the CC mixture
respectively. The creep coefficient in fibre reinforced SCC mixtures (D-SCC, S-SCC and
DS-SCC) is lower than those of the EPS-LWC and N-SCC mixtures. However, the creep
coefficient in D-SCC and S-SCC mixtures is higher than that of CC mixture. The final
creep coefficient in DS-SCC is very close to that of CC mixture.

266
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

6.9. SUMMARY

The experimental program of the current study is divided into three major parts. The
first part illustrated in this chapter, deals with the mechanical properties of fresh and
hardened lightweight concrete made with EPS beads to manufacture the testing samples
and slab specimens. Number and type of the required tests and the relevant Australian
and international standards used to define the required properties of the concrete are
given also.
The particular mixture design to achieve the required density and compressive strength
is presented and the technical information of the cement, fine and coarse aggregates and
admixture are also given in detail.
Preparation methods and curing conditions before testing the samples are illustrated.
The hardened properties of concrete including compressive strength, splitting tensile
strength, modulus of elasticity and modulus of rupture and pull out test are performed at
the ages of 3, 7, 14, 21, 28, 56 and 91 days. Moreover, steel concrete bond properties,
compressive stress-strain curve and creep and shrinkage of concrete are measured.
The test ages and the shape and size of the test specimens for hardened LWC are
shown in Table (6.32). Three LWC specimens have been tested per age for each
property.

Table 6.32: Shape and size of the LWC test specimens and test ages

Cylinder Cube Prism Test ages


Property
(mm) (mm) (mm) (days)
Compressive strength 75×150, 100×200,150×300 100, 150 3,7, 14, 21, 28, 56, 91
Modulus of elasticity 150×300 3,7, 14, 21, 28, 56, 91
Modulus of rupture 100×100×400 3,7, 14, 21, 28, 56, 91
Splitting tensile
100×200,150×300 3,7, 14, 21, 28, 56, 91
strength
Steel-concrete Bond 150×300 3,7, 14, 21, 28, 56, 91
Pull-out test 150×300 7, 14, 28, 56, 91
Stress-strain 3,7, 14, 21, 28, 56, 91
Creep 150×300 Up to 210 days
Shrinkage 75×75×280 Up to 450 days

Table (6.33) explains the applied Australian standards and international codes for each
test in the laboratory conditions.

267
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

Table 6.33:Applied codes on the curing, sampling and testing the LWC specimens

Property Test method Property Test method

AS 1012.8.1 (2014) , AS 1012.8.2 (2014),


Compressive strength Modulus of rupture
ASTAM C39 (2000) ASTM C1018 (2002)

AS 1012.8.1(2014),
Splitting Tensile strength Steel-concrete Bond RILEM –RC6 (2004)
ASTM C496 (2000)

Creep AS 1012.16 (2014) Pull-out test RILEM –RC6 (2004)

AS 1012.17 (2014),
Shrinkage AS-1012.13 (2014) Modulus of elasticity
ASTM C469 (2002)

Table (6.34) explains the applied Australian standards AS-1012 (2014) to curing,
sampling, slump test and mass per volume measurements of the fresh LWC.

Table 6.34:Test methods to determine the fresh properties of lightweight oconcrete

Curing Sampling Slump test Mass per volume

Test method AS-1012.8 (2014) AS-1012.1 (2014) AS1012.3.1 (2014) AS1012.5 (2014)

Table (6.35) shows the mixture design of EPS lightweight concrete in this study.

Table 6.35: Mix design of lightweight concrete containing expanded polystyrene beads (SSD condition)

Component Proportion
Cement (kg/m3) 500
3
Water (litre/ m ) 180
Water to cement ratio (w/c) 0.36
3
BST aggregate (Litre/ m ) 300
Coarse sand (kg/ m3) 310
Fine aggregate
Fine sand (kg/ m3) 310
3
Coarse aggregate (kg/ m ) 800
3
Admixture (litre/ m ) 2
3
Mixture volume (m ) 1.3
Slump (mm) 130
3
Fresh density (kg/m ) 2000

Table (6.36) summarizes the mechanical properties of EPS-LWC.


268
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

Table 6.36: Mechanical properties of EPS-LWC at different ages

Age (day) 3 7 14 21 28 56 91

Compressive strength (MPa) 14.49 25.38 27.93 29.62 30.96 32.33 33.3

Splitting tensile strength (MPa) 1.44 2.3 2.6 2.78 2.89 2.98 3.08

Modulus of Elasticity (MPa) 11.63 15.68 17.65 19.48 20.32 21.1 21.89

Modulus of rupture (MPa) 2.982 3.498 4.056 4.218 4.305 4.374 4.524

- Time-dependent mechanical properties of lightweight concrete

The following equations are developed in this study to explain the relationship between
each mechanical property of EPS-LWC at different ages with 28 days.

Parameter Equation

Compressive strength (MPa, days) ݂௖ᇱ ሺ‫ݐ‬ሻ ൌ ሺͲǤͳ͸͵͵ Ž ‫ ݐ‬൅ ͲǤͶͳ͹ሻ ݂௖ିଶ଼ 

Splitting tensile strength (MPa, days) ݂௖௧ ሺ‫ݐ‬ሻ ൌ ሺͲǤͳͷ͸ Ž ‫ ݐ‬൅ ͲǤͶ͵Ͷͳሻ ݂௖௧ିଶ଼

Modulus of elasticity (GPa, days) ‫ܧ‬௖ ሺ‫ݐ‬ሻ ൌ ሺͲǤͳͶ͸͹ Ž ‫ ݐ‬൅ ͲǤͶ͸͸͸ሻ ‫ܧ‬௖ିଶ଼

Modulus of rupture (MPa, days) ݂௥ ሺ‫ݐ‬ሻ ൌ ሺͲǤͳͲͶ͸ Ž ‫ ݐ‬൅ ͲǤ͸ʹሻ ݂௥ିଶ଼

- Shrinkage strain

The shrinkage strain of EPS lightweight concrete is studied under different curing,
drying and rewetting conditions. The ultimate shrinkage strain of EPS lightweight
concrete after 450 days is considerably higher than that of the conventional and self-
compacting concrete in the same grade of compressive strength.
Different conditions of initial curing, drying and rewetting were applied to the
specimens as reported in Figure (6.60). The relative humidity and temperature of the
environment were kept at 50%±5 and 23 ±2oC respectively.

269
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

Figure 6.60: Initial curing, drying and rewetting conditions of shrinkage prisms

Among the existing models of predicting the shrinkage strain of lightweight concrete,
the model proposed by Best and Polivka (1959), gives the most compatible results with
the experimental data. Therefore, it is selected as the basis of the model proposed in this
study. Figure (6.61) shows compatibility of the proposed model with the experimental
data.

1400

1200

1000
shrinkage data (με

800

600

400

ACI-435-R95 AS-3600(2009)
200 ACI-209(2008) Best and Polivka (1959)
Experimental data Proposed model

0
0 50 100 150 200 250 300 350 400 450
Age after 7 days moist curing (day)

Figure 6.61: Experimental data of EPS-LWC versus predicted shrinkage strain by different models

The following models are proposed in this study to predict the shrinkage strain of
lightweight concrete; also an accurate relationship is established between the rate of
weight loss and the age of shrinkage specimens.

Parameter Equation
Shrinkage strain (με, days) ߝ௦௛ ൌ ͳͺͺ ൈ ‫ ݐ‬଴Ǥଷଵହ
Rate of weight loss (gr/days, days) ܸοௐ ൌ ͶͳǤͻ͵͹ ൈ ‫ ݐ‬଴Ǥ଻ଽ଼

270
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

- Creep coefficient

Corresponding to two different levels of sustained loads on the slab specimens in the
instantaneous and time-dependent deflection, two groups of creep specimens were loaded
at the age of 14 days. The creep specimens were loaded to 30% and 40% of the ultimate
compressive strength of the used EPS-LWC mixture. Each pair of creep rigs consisted of
2 cylinders.

The final creep coefficient after 225 days under loading levels 1 and 2 were 2.231 and
2.415 respectively. The maximum creep coefficient of specimens subjected to loading
level 2 is 8.2% higher than that of loading level 1. Figure (6.62) evidently shows the
effect of the loading level on the creep coefficients.

2.4
2.2
2
1.8
1.6
Creep coefficient

1.4
1.2
LOAD-30%
1
LOAD-40%
0.8
0.6
0.4
0.2
0
0 25 50 75 100 125 150 175 200 225 250
Age (day)

Figure 6.62: Creep coefficients of LWC specimens under sustained loads of 30% and 40% of the
ultimate capacity

Figure (6.63) shows the ratio of creep coefficients at different ages in two creep rigs,
compared to the constant ratio of loading levels. Despite the constant ratio of applied
loads during the test in two creep rigs, the ratio of creep coefficients is following a
descending rate from 25% at first days to 8% at age 225 days. Similar to descending trend
of creep coefficients, the ratio of creep strain is following a descending rate from 53% at
first days to 37% at the age of 225 days.

The final strain ratio is 27% higher than the ratio of applied loads in two rigs.
However, the final creep coefficient ratio is close to the ratio of applied loads. In other
words, the ratio of the final creep coefficient after 225 days is more compatible with the
ration of applied load levels.

271
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

25
Ratio of creep coefficients

20 Ratio of load levels

15

Ratio (%)
10

0
0 50 100 150 200 250
Age (day)
Figure 6.63: Comparing the decreasing ratio of creep coefficient at different ages

Equation (6.21) is proposed in this study to predict the creep strain of EPS-LWC
cylindrical specimen respectively. Effect of the magnitude of the applied sustained stress
on the creep specimen is also included in the models.

ߝ௖௥ ൌ ͺͲͲߙ‫ݐ‬ఉ Eq. (6. 21)

Where; ߝ௖௥ is creep strain, α is the ratio of applied stress to the compressive strength of
concrete, β is the coefficient of time.

Sustained stress (MPa) ߙ β

0.3݂௖ᇱ 0.3 00.375

0.4݂௖ᇱ 0.4 0.375

Equations (6.22) is proposed in this study to predict the creep coefficient of EPS-LWC
cylindrical specimen respectively

௧ೌ
‫׎‬ሺ‫ݐ‬ǡ ‫ݐ‬଴ ሻ ൌ ‫׎‬ Eq. (6. 22)
௕ା௧ ೎ ௨

Where; ‫׎‬ሺ‫ݐ‬ǡ ‫ݐ‬଴ ሻ is creep coefficient, ‫ ࢛׎‬is ultimate creep coefficient, and a, b, c are
experimental parameters.

272
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

The predictions of these models are verified by the experimental data of creep strain
and creep coefficient at two different levels of sustained compression.

Sustained stress (MPa) a b c

0.3݂௖ᇱ 0.65 8 0.6

0.4݂௖ᇱ 0.75 10 0.72

- Pull-out test

The pull-out test is performed to measure the bond between reinforcement and
concrete. Table (6.37) describes the measured values of bond and slip between the
concrete and the reinforcement at different ages.

Table 6.37: Bond stress and corresponding slip values of test specimens at different

Age (days) 3 7 14 21 28 56 91

Bond (N/mm2) 14.83 17.77 19.33 20.16 21.08 22.87 23.73

Slip (mm) 33.3 25.3 21.3 17.7 15 11.3 12.3

The following model in Equation (6.23) is developed in this study to describe the
relationship between the bond-stress and the slip in a reinforced lightweight concrete
member under direct tension.

߬ ൌ ͸ʹǤͻʹͺ‫ି ݏ‬଴Ǥଷଽଽ  Eq. (6. 23)

Where; τ is bond stress (MPa) and s is the slip of steel bar under tension (mm).

- Compressive stress-strain curve

The stress-strain behaviour of the lightweight concrete cylinder specimens under


compression at two ages of 14 and 28 days are given in Figure (6.64) as samples of all
tests. For each age, three cylinder specimens were tested and considering the good
reproduction of test results, the final stress-strain curves for each age were taken as the
average of three curves.

273
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

35
14 days
30
28 days

Compressive stress (Mpa)


25

20

15

10

0
0 0.001 0.002 0.003 0.004 0.005
Strain

Figure 6.64: Stress-strain diagram of the LWC mixture at different ages

Table (6.38) shows absorbed energy per unit volume of cylinder specimens under
compression (Gc) values at different ages.

Table 6.38: The energy absorbed in cylinder specimens under compression

Age (days)
3 7 14 21 28 56 91
Gc (MPa) 0.02185 0.03892 0.04735 0.05477 0.0632 0.06856 0.07732

Equation (6.24) is developed to describe the development of absorbed energy by age.

‫ܩ‬௖ ሺ‫ݐ‬ሻ ൌ ሺͲǤͲͳ͸ Ž ‫ ݐ‬൅ ͲǤͲͲ͸ͳሻ ‫ܩ‬௖ିଶ଼ Eq. (6. 24)

Where; ‫ܩ‬௖ ሺ‫ݐ‬ሻ is the stored energy per unit volume of the specimen (MPa) at the age of
t (days) and ‫ܩ‬௖ିଶ଼ is stored energy per unit volume of the specimen (MPa) at the age 28
days.

Variation of stored energy per unit volume under compression with the strain changes
at different ages is shown in Figure (6.65).

274
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

0.09
3 days
0.08
7 days

Stored energy per unitt volume (MPa)


0.07 14 days
0.06 21 days
28 days
0.05
56 days
0.04 91 days
0.03

0.02

0.01

0
0 0.001 0.002 0.003 0.004 0.005
Strain (mm/mm)

Figure 6.65: Stored energy under compression versus strain of EPS-LWC at different ages

- Relationships between compressive strength and mechanical properties

The compressive strength can illustrate mechanical properties of hardened concrete.


The following equations have been developed in this study to describe the relationship
between the compressive strength and other mechanical properties of concrete.

Parameter Equation

Compressive strength (MPa) ݂௖ᇱ ሺ‫ݐ‬ሻ ൌ ሺͲǤͳ͸͵͵ Ž ‫ ݐ‬൅ ͲǤͶͳ͹ሻ ݂௖ିଶ଼ 

Splitting tensile strength (MPa) ଴Ǥଽଵ଻଺


݂௖௧ ൌ ͲǤͳʹʹ͹ ൈ ݂௖ᇱ

Modulus of elasticity (GPa, MPa) ଴Ǥ଻ହଶଷ


‫ܧ‬௖ ൌ ͳǤͷͲͶͷ ൈ ݂௖ᇱ

Modulus of rupture (MPa) ଴Ǥହ଴଴ଷ


݂௥ ൌ ͲǤ͹͸ʹͷ ൈ ݂௖ᇱ

Ultimate pull-out load (KN, MPa) ‫ ܨ‬ൌ ͲǤͻͻʹͷ݂௖ᇱ ൅ ͳ͹ǤͶͺͳ

- Comparing the mechanical properties of lightweight, self-compacting and


conventional concrete

The experimental results of this study are compared to the previously conducted
experiments on self-compacting concrete slabs by Aslani (2014) and conventional
concrete slabs by Nejadi (2005).

275
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

Table (6.39) explains the mixture proportions and density of LWC and SCC mixtures.

Table 6.39: Mixture design of self-compacting and lightweight concrete

Constituents N-SCC D-SCC S-SCC DS-SCC LWC CC


3
Cement (kg/m ) 160 160 160 160 250
3
Fly Ash (kg/m ) 130 130 130 130
3
GGBFS (kg/m ) 110 110 110 110
3
Cementitious content (kg/m ) 400 400 400 400
Water (kg) 208 208 208 208 150 lit/m3
Water Cementitious ratio 0.52 0.52 0.52 0.52 0.6
3
Aggregate (kg/m )
Coarse sand 660 660 660 660 620
fine sand 221 221 221 221
Coarse aggregate 820 820 820 820 800
3
Light Aggregate (BST) (lit/m ) 300
3
Admixtures (lit/m )
Super plasticizer 4 4.86 4.73 4.5
Viscosity modifying agent 1.3 1.3 1.3 1.3
High range water reducing 1.6 1.6 1.6 1.6 2
agent
Fibre content (kg/m3)
Steel - 30 - 15
Polypropylene - - 5 3
Density 2340 2274 2330 2385 2000 2450

Table (6.40) shows the measured compressive strength and modulus of elasticity of
EPS-LWC, SCC and CC mixtures at 14 and 28 days.

Table 6.40: Mechanical properties of SCC, LWC and CC mixtures

Property Age (day) N-SCC D-SCC S-SCC DS-SCC LWC CC


݂௖ᇱ ሺ‫ܽܲܯ‬ሻ 14 29.05 34.3 32.45 38.1 27.93 18.3
28 33.3 38 38.1 45 30.96 24.8
‫ܧ‬௖ ሺ‫ܽܲܩ‬ሻ 14 32.24 29.14 29.68 31.26 17.65 22.82
28 35.39 35.76 35.76 36.1 20.32 24.95

Figures (6.66) and (6.67) compare the measured values of compressive strength and
modulus of elasticity of the mixtures at ages 14 and 28 days respectively.

276
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

50
14 days 28 days
45

Compressive strength (MPa)


40

35

30

25

20

15
N-SCC D-SCC S-SCC DS-SCC LWC NC
Mixtures

Figure 6.66: Compressive strength of EPS-LWC, SCC and CC mixtures at 14 and 28 days

40
14 days 28 days
Modulus of Elasticity (GPa)

35

30

25

20

15
N-SCC D-SCC S-SCC DS-SCC LWC NC
Mixtures

Figure 6.67: Modulus of elasticity of EPS-LWC,SCC and CC mixtures at 14 and 28 days

Figure (6.68) compares the modulus of rupture (flexural tensile strength) of CC, LWC
and SCC mixtures at different ages.

277
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

Modulus of rupture (MPa)


6

3
N-SCC D-SCC
2
S-SCC DS-SCC
1 LWC CC

0
0 20 40 60 80 100
Age (day)

Figure 6.68: Comparison of the modulus of rupture in LWC, CC and SCC mixtures

According to Figure (6.68), EPS-LWC poses the lowest values of the modulus of
rupture, especially after age 28 days. However, SCC mixture with and without fibre
reinforcing are in the top range of the modulus of rupture at almost all ages.

- Shrinkage strain

Figure (6.69) compares the shrinkage strain of EPS-LWC, SCC and CC mixtures up to
240 days.

1200

1000
Shrinkage strain (με)

800

600
LWC
400 CC
N-SCC
D-SCC
200
DS-SCC
S-SCC
0
0 50 100 150 200 250
Age (days)
Figure 69: Comparison of shrinkage strain in EPS-LWC, SCC and CC mixtures
278
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

According to the Figure (6.69), the shrinkage strain in LWC is considerably higher
than that of the CC and SCC mixtures. The free shrinkage strain in LWC and N-SCC
mixtures is 51% and 10% greater than that of the CC mixture. The porous (permeable)
nature of the lightweight concrete, relatively higher ratio of water to cement and high rate
of the water loss cause higher values of the drying shrinkage in lightweight concrete.

Figure (6.70) compares the creep coefficient in EPS-LWC, SCC and CC specimens
under sustained compressive stress. The specimens were loaded at age 14 days. The
applied stress is equal to 40% of the compressive strength of the companion cylinders in
each type of concrete.

2.5

2
Creep coefficient

1.5

CC-40%
1
LWC-40%
N-SCC-40%

0.5 D-SCC-40%
S-SCC-40%
DS-SCC-40%
0
0 50 100 150 200 250
Age (day)
Figure 6.70: Creep coefficient of CC, SCC and EPS-LWC specimens loaded at age 14 days

According to Figure (6.70) the creep coefficient in EPS-LWC is considerably higher


than that of SCC and CC specimens under the same level of sustained compressive stress.

The creep coefficient in EPS-LWC and CC are the highest and lowest recorded values
after 240 days. The creep coefficient at 240 days in LWC and N-SCC mixtures is 1.61
and 1.3 times higher than that of the CC mixture respectively. The creep coefficient in
fibre-reinforced SCC mixtures (D-SCC, S-SCC and DS-SCC) is lower than those of the
EPS-LWC and N-SCC mixtures. However, the creep coefficient in the D-SCC and S-
SCC mixtures is higher than that of the CC mixture. The final creep coefficient in DS-
SCC is very close to that of the CC mixture.

279
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

REFERENCES

1. Aslani, F., Experimental and numerical study of time-dependent behaviour of reinforced


self-compacting concrete slabs. PhD dissertation, University of Technology Sydney
(UTS), 2014.
2. Nejadi, S., Time-dependent cracking and crack control in reinforced concrete structures.
2005, PhD thesis, The University of New South Wales Sydney, Australia.
3. AS-1411, Methods for sampling and testing aggregates, particle size distribution-sieving
method. Standards Australia, Sydney, Australia, 2011.
4. RTA-06, Materials test methods. Regianl transportation authority , Chichago, IL, USA.,
2006. 1.
5. AS-1141.11.1, Methods for sampling and testing aggregates - Particle size distribution -
Sieving method. Standards Australia, Sydney, Australia, 2009.
6. AS-1141.12, Methods for sampling and testing aggregates - Materials finer than 75 um
in aggregates (by washing). Standards Australia, Sydney, Australia, 2011.
7. AS-1141.14, Methods for sampling and testing aggregates - Particle shape, by
proportional caliper. Standards Australia, Sydney, Australia, 2011.
8. AS-1141, Methods for sampling and testing aggregates Standards Australia, Sydney,
Australia, 2011.
9. AS-1141.4, Methods for sampling and testing aggregates - Bulk density of aggregate.
Standards Australia, Sydney, Australia, 2011.
10. AS-1141.6.1, Methods for sampling and testing aggregates - Particle density and water
absorption of coarse aggregate - Weighing-in-water method. Standards Australia,
Sydney, Australia, 2011.
11. AS-1141.22, Methods for sampling and testing aggregates - Wet/dry strength variation.
Standards Australia, Sydney, Australia, 2008.
12. AS-1141.23, Methods for sampling and testing aggregates - Los Angeles value.
Standards Australia, Sydney, Australia, 2009.
13. AS-1141-24, thods for sampling and testing aggregates - Aggregate soundness -
Evaluation by exposure to sodium sulphate solution. Standards Australia, Sydney,
Australia, 2013.
14. AS-1141.32, Methods for sampling and testing aggregates - Weak particles (including
clay lumps, soft and friable particles) in coarse aggregates. Standards Australia, Sydney,
Australia, 2008.
15. AS-1141.25.2, Methods for sampling and testing aggregates - Degradation factor -
Coarse aggregate. Standards Australia, Sydney, Australia, 2013.
16. AS-1141.13, Methods for sampling and testing aggregates - Material finer than 2
micrometer. Standards Australia, Sydney, Australia, 2007.
17. AS-1141.34, Methods for sampling and testing aggregates - Organic impurities other
than sugar. Standards Australia, Sydney, Australia, 2007.
18. AS-1141.35, Methods for sampling and testing aggregates - Sugar. Standards Australia,
Sydney, Australia, 2007.
19. AS-1478.1, Methods for sampling and testing aggregates, particle size distribution-
sieving method. Standards Australia, Sydney, Australia, 2000.
20. AS-3972, General purpose and blended cements. Standards Australia, Sydney, Australia,
2010.
21. AS-2350, Methods of testing Portland and blended cements. Standards Australia,
Sydney, Australia, 2006.

280
Chapter 6: Experimental program- phase I, material properties of lightweight concrete
containing expanded polystyrene beads

22. AS-2350.2, Methods of testing portland, blended and masonry cements - Chemical
composition. Australian Standard, 2006.
23. AS-2350.8, Methods of testing portland, blended and masonry cements - Fineness index
by air permeability method. Australian standard, 2006.
24. AS-2350.3, Methods of testing portland, blended and masonry cements - Normal
consistency. Standards Australia, Sydney, Australia, 2006.
25. AS-2350.13, Methods of testing portland, blended and masonry cements - Determination
of drying shrinkage of cement mortars. Standards Australia, Sydney, Australia, 2006.
26. AS-2350.11, Methods of testing portland, blended and masonry cements - Compressive
strength. Standards Australia, Sydney, Australia, 2006.
27. AS-2350.4, Methods of testing portland, blended and masonry cements - Setting time.
Standards Australia, Sydney, Australia, 2006.
28. AS-2350-5, Methods of testing portland, blended and masonry cements - Determination
of soundness. Australian Standard, 2006.
29. AS-1210.8.1, Methods of testing concrete- Method 8.1: method for making and curing
concrete, compression and indirect tensile test specimens. Standards Australia, Sydney,
Australia, 2014.
30. AS-1012, Methods of testing concrete,. Standards Australia, Sydney, Australia, 2014.
31. AS-3600-09, Concrete structures. Australian Standard, 2009.
32. Best, C.H. and M. Polivka Creep of lightweight concrete. Magazine of Concrete
Research, 1959. 11, 129-134.
33. ACI-209.2R-08, Guide for Modeling and Calculating Shrinkage and Creep in hardened
concrete. American Concrete Institute (ACI), ACI Committee 209, Farmington Hills, MI
48331, U.S.A, 2008.
34. ACI-435-R95, Control of Deflection in Concrete Structures. American Concrete Institute
(ACI) Committee 435, 2003.
35. Neville, A.M., Properties of Concrete. 4th ed , Birmingham, UK, 2010.

281
Chapter 7 – Flexural load-deflection behaviour

CHAPTER 7

EXPERIMENTAL PROGRAM (PHASE II) –


FLEXURAL LOAD-DEFLECTION BEHAVIOUR
Chapter 7 – Flexural load-deflection behaviour

Table of Contents

7.1. SHORT-TERM EXPERIMENTAL PROGRAM ................................................................... 282


7.2. TEST PARAMETERS AND REINFORCEMENT LAYOUT .............................................. 283
7.3. CONSTRUCTION OF THE SPECIMENS AND TEST PROCEDURE IN SHORT-TERM
LOAD-DEFLECTION TEST .................................................................................................................. 284
7.4. LOADING BLOCKS ............................................................................................................. 285
7.5. INSTRUMENTATION OF THE SLABSAND TEST SETUP .............................................. 286
7.5.1. Strain Gauges ..................................................................................................................... 286
7.5.2. DEFLECTION RECORDING DEVICE ............................................................................ 287
7.6. LOAD-DEFLECTION BEHAVIOUR ................................................................................... 288
7.6.1. Test Results of Lightweight Concrete Slabs ....................................................................... 289
7.6.2. Material Properties: ............................................................................................................ 290
7.6.3. Load-deflection behaviour of slab LWC-1 ......................................................................... 291
7.6.4. Load-deflection behaviour of slab LWC-2 ......................................................................... 292
7.6.5. Load-deflection behaviour of slab LWC-3 ......................................................................... 293
7.6.6. Load-deflection behaviour of slab LWC-4 ......................................................................... 295
7.6.7. Comparison of load-deflection curves in all LWC slabs .................................................... 296
7.7. PREVIOUS EXPERIMENTS ON LOAD-DEFLECTION BEHAVIOUR OF SLABS ........ 297
7.7.1. Short-term loading of conventional concrete slabs (Nejadi, 2005) .................................... 297
7.7.2. Short-term loading of self-compacting concrete slabs (Aslani, 2014) ............................... 300
7.8. COMPARISON OF LOAD-DEFLECTION CURVES IN LIGHTWEIGHT,
CONVENTIOANL AND SELF-COMAPCTING CONCRETE SLABS................................................ 302
7.9. SUMMARY............................................................................................................................ 304
REFERENCES ......................................................................................................................................... 308
Chapter 7 – Flexural load-deflection behaviour

List of Figures
Figure 7.1: Slab moulds before and after pouring of lightweight concrete .............................................. 285
Figure 7.2: Dimensions and reinforcement details for LWC slab specimens .......................................... 285
Figure 7.3: Loading blocks and arrangement to supply two loading levels on slabs ............................... 286
Figure 7.4: a) arrangement of strain gauges attached to steel bars and b) Position and number of the strain
gauges on concrete surface (side and top) ....................................................................................... 287
Figure 7.5: Installed strain gauges on steel bars and surface of concrete ................................................. 287
Figure 7.6: Installation of LVDT at two points of slab span .................................................................... 288
Figure 7.7: Load- deflection test arrangements for all specimens ............................................................ 289
Figure 7.8: Load-deflection test setup and slab failure ............................................................................ 290
Figure 7.9: Deflection and cracking pattern of LWC-1 at failure point .................................................. 292
Figure 7.10: Load-deflection curve of slab LWC-1 at mid-span.............................................................. 292
Figure 7.11: Deflection and cracking pattern of LWC-2 at failure point ................................................. 293
Figure 7.12: Load-deflection curve of slab LWC-2 at mid-span.............................................................. 293
Figure 7.13: Deflection and cracking pattern of LWC-3 at failure point ................................................. 294
Figure 7.14: Load-deflection curve of slab LWC-3 at mid-span.............................................................. 294
Figure 7.15: Deflection and cracking pattern of LWC-4 at failure point ................................................. 295
Figure 7.16: Load-deflection curve of slab LWC-4 at mid-span.............................................................. 296
Figure 7.17: Comparing the load-deflection curves in four LWC slabs ................................................... 296
Figure 7.18: Test arrangement for all the beams and slab in short-term loading [1] ................................ 298
Figure 7.19: Failure of a) slab S2-b and b) beam B1-a [1] ....................................................................... 299
Figure 7.20: Short-term load-defection curves for conventional concrete slabs ...................................... 299
Figure 7.21: General view of the loading cells, concrete strain gauges and LVDT set-up [2] ................. 301
Figure 7.22: Load-deflection curve for slab SCC and FRSCC slabs at mid-span .................................... 302
Figure 7.23: Comparison of the load-deflection curves in all LWC, SCC and CC slabs ......................... 302
Figure 7.24: Load-deflection curves in identical LWC, SCC and CC slabs ............................................ 303
Figure 7.25: Comparison of the load-deflection curves in LWC, SCC and CC slabs .............................. 304
Figure 10.26: Comparing the load-deflection curves in four LWC slabs ................................................. 306
Figure 7.27: Comparison of the load-deflection curves in EPS-LWC, SCC and CC slabs ...................... 307
Chapter 7 – Flexural load-deflection behaviour

List of Tables
Table 7.1: Details of slabs for short-term load-deflection testin of slabs ................................................. 285
Table 7.2: Mechanical properties of EPS-lightweight concrete used in the slab specimens .................... 291
Table 7.3: Detail of test specimens in the experiment by Nejadi (2005) [1] ............................................ 298
Table 7.4: Summary of the results from short-term flexural test [1] ....................................................... 299
Table 7.5: Detail of test specimens in the experiment by Aslani (2014) [2] ............................................ 300
Table 7.6: Summary of the results from short-term flexural test ............................................................. 301
Table 7.7: Failure load and maximum deflection of lightweight concrete slabs ...................................... 305
Table 7.8: Summary of the results from short-term flexural test ............................................................. 306
Chapter 7 – Flexural load-deflection behaviour

Chapter 7: Flexural load-deflection behaviour

7.1. SHORT-TERM EXPERIMENTAL PROGRAM

The experimental part of this study consists of the laboratory investigation of Light-Weight
Concrete (LWC) slabs to investigate the short-term load-deflection of slabs up to the failure
point, instantaneous deflection under tow different levels of service loads and long-term
defection under two levels of sustained service loads and the effects of creep and shrinkage.

Following are the main investigated parts in the experimental phase of this study:

a) Load-deflection behaviour of slab under short-term loading until failure;


b) Instantaneous flexural deflection of slabs under service loads;
c) Time-dependent flexural deflection of slabs under service load and effects of creep and
shrinkage;
d) Comparing the experimental results with the previously conducted laboratory
experiments on idem slabs made with Self-Compacting Concrete (SCC) and
Conventional Concrete (CC).

For instantaneous and time-dependent deflection study, different points and regions were
monitored to record the changes of deflection and strain in the slabs. Deflection was monitored
at mid-span and the point on the border of high moment region of the span length. Variation of
the strain in the tensile reinforcing bars embedded in the concrete was monitored within the high
moment region of the slab length. Strain of the concrete under service sustained loading and
effects of creep and shrinkage were recorded in the concrete surface at the level of

282
Chapter 7 – Flexural load-deflection behaviour

reinforcement steel. Additionally, the compressive strain of concrete on top surface at the mid-
span of slab was measured under the sustained loading and effects of creep and shrinkage.

For short-term load-deflection investigation, the strain and deflection were monitored at the
same positions under the load increment until the failure of the slabs.

Properties of concrete including the compressive strength, modulus of elasticity and modulus
of rupture and splitting tensile strength at different ages were measured on companion
specimens (as presented in Chapter Five).

The main objectives of the laboratory tests in the experimental phase are as follows:

- Better understanding of the mechanisms associated with LWC flexural deflection;


- To obtain benchmark, laboratory-controlled data to assist in the development of rational
design-oriented procedures for the control of deflection in slabs;
- To study the difference between elastic and instantaneous deflection of LWC slabs;
- To modify the ratio of long-term to short-term deflection of LWC slabs;
- Investigating the effect of loading history on deflection and strain changes on identical
specimens made of LWC;
- Proposing a new model for effective moment of inertia for defection calculations;
- To verify the accuracy of the existing models of LWC properties to simulate the
nonlinear behaviour of RC structures in Finite Element (FE) analysis in the ATENA
program;
- Proposing a new model for tensile stress-strain behaviour of the LWC and verifying it by
test results and FE analysis;
- To investigate the accuracy of the commonly used coefficient of creep and shrinkage (kcs)
and other existing models in the calculations by applying test results in FE analysis.

7.2. TEST PARAMETERS AND REINFORCEMENT LAYOUT

In addition to the above-mentioned objectives, instantaneous and long-term deflection and


short-term load-deflection behaviour are compared with the previously conducted experiments
on similar section and length of slabs made in other types of concrete. In this regard, the
experimental investigation by Nejadi (2005) [1] on the Conventional Concrete (CC) slabs and
the experimental investigation by Aslani (2014) [2] on the Self-Compacting Concrete (SCC)
slabs with and without fibre reinforcing have been considered for comparison and verification
purposes. Nejadi (2005) [1] investigated the time-dependent cracking of two slabs made of the

283
Chapter 7 – Flexural load-deflection behaviour

same batch of concrete, while Aslani (2014) [2] studied four pairs of one-way slabs made of 4
different batches of normal and fibre-reinforced SCC mixtures.

All LWC, CC and SCC slabs were singly reinforced by 4N12 steel bars at the tensile zone.
There was no shear reinforcement applied in the slabs. All slabs were simply supported with the
same span length. The applied sustained loads on the CC slabs were uniformly distributed loads
of about 20%, 30% of the ultimate bending capacity of the slab section. While, the sustained
loading level on the SCC and LWC slabs were about 30% and 40% of the ultimate bending
capacity of the slab section.

Overall, the changing parameters in the slabs of LWC, SCC and CC types are of mechanical
properties of concrete, shrinkage, creep and loading level. These parameters are investigated
and compared with the obtained results from the previously conducted experimental results on
idem slabs made with conventional and self-compacting concrete. Considering the LWC, CC
and SCC slabs, the parameters of tests varied, including one LWC mixture, one CC mixture and
four SCC mixtures.

This chapter explains the testing setup, investigated parameters and results of the short-term
load-deflection of slabs. The instantaneous and long-term deflections of slabs are
comprehensively discussed in Chapters 7 and 8 of this study, respectively.

7.3. CONSTRUCTION OF THE SPECIMENS AND TEST PROCEDURE IN


SHORT-TERM LOAD-DEFLECTION TEST

To investigate the short-term load-deflection of slabs made of Light-Weight Concrete


Containing Expanded Poly-Styrene beads (EPS-LWC), four simply-supported one-way slabs
were subjected to increasing loading up to the failure of the span.

All specimens were constructed in 3800 mm long formwork and were simply supported over
a 3500 mm span. Variations of the mechanical properties of concrete were measured on
companion cylinders and prisms at ages 3, 7, 14, 21, 28, 56 and 91 days.

Before casting each specimen, the inside surface of the mould was cleaned and coated with a
thin layer of concrete release agent to prevent adhesion of the concrete. The LWC was placed
into the mould in equal layers and compacted by internal vibration. For each layer, the concrete
was vibrated until the surface became smooth in appearance. Floating of the BST beads to the
concrete surface due to vibration also controlled the vibration duration of each layer. Sufficient
concrete was placed into the top layer to overfill the mould, after which the surface was stripped
off and finished with a steel trowel. The companion specimens were also cast at the same time

284
Chapter 7 – Flexural load-deflection behaviour

as the test specimens. Within two hours of casting, the slab specimens were covered with wet
hessian and plastic sheets and left in their moulds for three days. After three days they were
removed and kept moist continuously using a thick covering of wet hessian until the age of
testing. Figure (7.1) shows the slab moulds before and after pouring of LWC.

The slab specimens were each nominally 3500 mm long by 400 mm wide. In all slabs, the
nominal distance from the soffit of the centroid of the main reinforcement was 25 mm. Each
slab was reinforced with 4N12 steel bars. Details of the cross section for all the LWC, SCC and
CC slabs are presented in Figure (7.2). Also, Table (7.1) shows the reinforcement details and the
ratio of tensile bars in the effective section area.

Figure 7.1: Slab moulds before and after pouring of lightweight concrete

Figure 7.2: Dimensions and reinforcement details for LWC slab specimens

Table 7.1: Details of slabs for short-term load-deflection testin of slabs

No. of Bar dia. Steel area


Specimen cb(mm) cs(mm) s(mm) ρ=Ast/bd
bars (mm) (mm2)
Slab - a 4N12 25 40 103 12 452 0.0083
Slab - b 4N12 25 40 103 12 452 0.0083

7.4. LOADING BLOCKS

To study the effect of loading history on the deflection behaviour of the specimens, two
different levels of service loads at the age of 14 days have been applied to the top surface of the
slabs. To distribute the applied loads on the slabs uniformly, timber pads with 100mm width

285
Chapter 7 – Flexural load-deflection behaviour

×30mm thickness and 450 mm length were used at equal intervals under each loading block.
The dimension of loading blocks and centre to centre spacing of loading pads (timber pad)
under loading blocks are shown in Figure (7.3). All dimensions are in millimetres.

Figure 7.3: Loading blocks and arrangement to supply two loading levels on slabs

7.5. INSTRUMENTATION OF THE SLABSAND TEST SETUP

7.5.1. Strain Gauges


Since the critical area for the flexural behaviour of the simply supported slab is in the middle
one-third of span length, strain changes and deflection variation were monitored in this region.
According to Figure (7.4) different strain gauges were attached to the tensile reinforcement to
record the strain variation in steel bars under short-term loading, sustained service loads and the
effects of creep and shrinkage.

To monitor the strain changes in tensile concrete, a different type of strain gauges were
attached to the concrete surface at the same level and direction of the monitored strain in steel
bars. It helps to compare the changes of strain in concrete and reinforcement at the same level in
the tensile zone. Furthermore, the bond characteristics of the steel bars and concrete could be
investigated by these strain gauges. To record the compression strain on concrete surface, two
strain gauges in the same direction of middle steel bars were attached to the top surface of the
concrete slab at mid-span. Figure (7.5) shows the arrangement and positioning of strain gauges
attached to steel bars and concrete surface in LWC slabs. The strain gauges arrangement in SCC
and CC slabs are similar except for the top surface strain gauges.
286
Chapter 7 – Flexural load-deflection behaviour

Figure 7.4: a) arrangement of strain gauges attached to steel bars and b) Position and number of the strain
gauges on concrete surface (side and top)

WFLA6-11 and PL60-11/3L are two different types of strain gauges used to measure the
strain in steel bars and concrete surfaces respectively. The WFLA-6-11 strain gauge is a type of
waterproof strain gauge having a pre-attached vinyl lead wire and a continuous coating with
epoxy resin. The coating is transparent and flexible, which ensures easy installation of the
gauge. PL-60-11 is a wire strain gauge having polyester resin backing. It is mainly used for
measurement of strain in concrete, mortar or rock. This gauge has a pre-attached vinyl lead wire
to a P series. Works for lead wire connection such as strain gauge terminal installation and lead
wire soldering are not required, which saves much time and labour. Figure (7.5) shows these
types of strain gauges applied in the current study.

Figure 7.5: Installed strain gauges on steel bars and surface of concrete

7.5.2. DEFLECTION RECORDING DEVICE


The deflection was recorded in the mid-span of slabs under short-term loading, instantaneous
and time-dependent sustained loading by a MTS-Temposonics R-Series position sensor.
287
Chapter 7 – Flexural load-deflection behaviour

Furthermore, to compare the deflection behaviour in high flexural moment regions, another
point on the border of the critical zone i.e. the point in one-third of span length was monitored.
The Linear Variable Differential Transformer (LVDT) (also called just a differential
transformer, linear variable displacement transformer, or linear variable displacement
transducer) was used to monitor the deflection. It is a type of electrical transformer used to
measure the linear displacement (position). Another version of this device that is used to
measure rotary displacement is called a Rotary Variable Differential Transformer (RVDT). The
LVDT transforms a position or linear displacement from a mechanical reference (zero, or null
position) into a relative electrical signal containing phase (for direction) and amplitude (for
distance) information. Figure (7.6) shows the installed LVDT at two points of each slab to
record the deflection variation.

Figure 7.6: Installation of LVDT at two points of slab span

7.6. LOAD-DEFLECTION BEHAVIOUR

For short-term investigation of load-deflection relationship, four singly reinforced LWC


slab specimens were cast and moist cured for 28 days. After 28 days the wet hessian was
removed and the specimens were identified and tested at different ages. Each specimen was
slowly and gradually loaded to failure over a period of approximately four hours. All the
specimens were simply supported on a 3.5 m span and tested to failure to investigate the
mid-span deflection, changes of strain in the critical bending region and distribution and
extent of primary and secondary cracking under short-term loading using two equal point
loads applied at the third points of the span, at ages greater than 28 days. Deflection at the
mid-span was monitored. The schematic diagram of the test set-up is shown in Figure (7.7).

288
Chapter 7 – Flexural load-deflection behaviour

Figure 7.7: Load- deflection test arrangements for all specimens

The position sensor and resistance strain gauge measurements were retrieved from a data
acquisition system connected to a computer while the load was supplied by a hydraulic jack,
attached to an electrically powered pressure pump. Figure (7.8) shows views of the
experimental set-up. Each specimen was simply supported at each end before testing, and then
one Temposonics R-Series position sensor was attached at the mid-span and linked to a
computer. Details of supports and the sensor are shown in Figure (7.9) also. Along with the
measured strain variation in steel bars, the changes of strain in the concrete surface at the level
of reinforcement and the top surface at the mid-span were recorded by different type of strain
gauges.

7.6.1. Test Results of Lightweight Concrete Slabs


To set the basis of the monitored data, and to calculate the exact values of deflection at mid-
span and the strain changes by increasing the loading value, the initial readings were taken at
zero load condition. To have an accurate estimation, the ultimate load capacity of the section
was calculated before the test. The load was then applied by 3 KN increments in each step on
the slab specimens. Although this study emphasises the deflection behaviour, every visible
crack on the tensile concrete surface was monitored at each load increment. This rate of loading
was maintained until about 60% of the expected ultimate load capacity. Then, the load was
gradually increased monotonically to failure and crack patterns were recorded at each load step.

289
Chapter 7 – Flexural load-deflection behaviour

Figure 7.8: Load-deflection test setup and slab failure

The experimental results from four lightweight concrete slabs made from the same concrete
mixture are presented. The results contains the measured material properties, variation of the
strain in eight points of the steel bars, and the strain in the concrete surface at the level of steel
bars and two point on top surface of the concrete. Also the load-deflection diagrams at the mid-
span of the slabs are presented. The results from each slab are discussed in detail and the results
from all slabs are also compared. The relevant diagrams for concrete strain at top and side
surfaces, steel strain, and some extra view of the test results are presented in Appendix-C.

7.6.2. Material Properties:


Table (7.2) gives a summary of the mechanical properties of the lightweight concrete used in
the slab specimens. Comprehensive information of the testing procedure and the detailed results
are given in Chapter 5 of this study. The mechanical properties are performed by standard test
methods and specimens to determine the compressive strength, splitting tensile strength,
modulus of rupture and modulus of elasticity at the ages of 3,7,14,21,28,56 and 91 days.

290
Chapter 7 – Flexural load-deflection behaviour

Table 7.2: Mechanical properties of EPS-lightweight concrete used in the slab specimens

Specimen age (days)


Property
3 7 14 21 28 56 91

Compressive strength (MPa) 14.49 25.38 27.93 29.62 30.96 32.33 33.3

Modulus of rupture (MPa) 1.44 2.3 2.6 2.78 2.89 2.98 3.08

Modulus of elasticity (GPa) 11.63 15.68 17.65 19.48 20.32 21.1 21.89

Splitting tensile strength (MPa) 2.982 3.498 4.056 4.218 4.305 4.374 4.524

7.6.3. Load-deflection behaviour of slab LWC-1


The first cracking at the mid-span took place approximately at load of 6 KN (3 KN from
each loading cell). Then, the increasing load increased the number and width of the cracks, and
caused the gradual growth in the mid-span deflection of up to 60% of the expected ultimate
load. The rate of deflection increment was higher at higher loading levels by increasing the
crack patterns and width. The majority of the crack pattern happened at the high moment region
due to tensile stress.

Figure (7.9) shows the crack pattern and deflection of slab LWC-1 at the failure point.

The measured deflection at mid-span versus applied load in the slab LWC-1 is shown in
Figure (7.10). The amount of tensile reinforcement in the slab provides an under-reinforced
section with a probability of ductile failure. The test showed a good ductile failure of slab LWC-
1 with an extended flat plateau in the load-deflection curve. The ultimate capacity was reached
when P=46.5 KN (total load of 2 load cells), when crushing of the top compressive surface
occurred. The maximum deflection corresponding to the failure point was 95.9 mm. The slab
LWC-1 failed in flexure in the pure moment zone, however, the first failure point with the
compressive concrete crushing above the crack was not exactly in the mid-span of slab.

291
Chapter 7 – Flexural load-deflection behaviour

Figure 7.9: Deflection and cracking pattern of LWC-1 at failure point

60

50

40
Load (KN)

30

20

10
LWC-1
0
0 20 40 60 80 100 120
Deflection (mm)

Figure 7.10: Load-deflection curve of slab LWC-1 at mid-span

7.6.4. Load-deflection behaviour of slab LWC-2


The first cracking at the mid-span took place approximately at load of 6 KN (3 KN from
each loading cell). Then, the increasing load increased the number and width of the cracks, and
caused the gradually growth in the mid-span deflection of up to 60% of the expected ultimate
load. The rate of deflection increment was higher at higher loading levels by increasing the
crack patterns and width.

Figure (7.11) shows the crack pattern and deflection of slab LWC-2 at the failure point.

292
Chapter 7 – Flexural load-deflection behaviour

Figure 7.11: Deflection and cracking pattern of LWC-2 at failure point

The measured deflection at mid-span versus applied load is shown in Figure (7.12). The test
showed a good ductile failure of the slab LWC-2 in the load-deflection curve. The ultimate
strength was recorded as P=48 KN, when the top compressive surface was crushed. The
maximum deflection of the slab at the failure point was 88 mm. The slab LWC-2 failed in
flexure in the pure moment zone, however, the first failure point with the compressive concrete
crushing above a crack was close to but not exactly at the mid-span of slab.

60

50

40
Load (KN)

30

20

10 LWC-2

0
0 20 40 60 80 100 120
Deflection (mm)

Figure 7.12: Load-deflection curve of slab LWC-2 at mid-span

7.6.5. Load-deflection behaviour of slab LWC-3


The first cracking at the mid-span took place approximately at load of 6 KN (3 KN from
each loading cell). Then, the increasing load increased the number and width of the cracks, and

293
Chapter 7 – Flexural load-deflection behaviour

caused the gradual growth in the mid-span deflection of up to 60% of the expected ultimate
load. The rate of the deflection increment was higher at higher loading levels by increasing the
crack patterns and width.

Figure (7.13) shows the crack pattern and deflection of slab LWC-3 at the failure point.

The measured deflection at mid-span versus applied load is shown in Figure (7.14). The test
showed a good ductile failure of the slab LWC-3 in the load-deflection curve. The ultimate
strength was recorded at P=41.53 KN, when the top compressive surface was crushed. The
maximum deflection of slab at the failure point was 114.7 mm. The slab LWC-3 failed in
flexure in the pure moment zone, however, the first failure point with the compressive concrete
crushing above a crack was close to but not exactly at the mid-span of slab.

Figure 7.13: Deflection and cracking pattern of LWC-3 at failure point

60

50

40
Load (KN)

30

20

10 LWC-3

0
0 20 40 60 80 100 120
Deflection (mm)

Figure 7.14: Load-deflection curve of slab LWC-3 at mid-span

294
Chapter 7 – Flexural load-deflection behaviour

7.6.6. Load-deflection behaviour of slab LWC-4


The first cracking at the mid-span took place at approximately load of 6 KN (3 KN from
each loading cell). Then, the increasing load increased the number and width of the cracks, and
caused the gradual growth in the mid-span deflection of up to 60% of the expected ultimate
load. The rate of deflection increment was higher at higher loading levels by increasing the
crack patterns and width.

Figure (7.15) shows the crack pattern and deflection of slab LWC-4 at the failure point.

Figure 7.15: Deflection and cracking pattern of LWC-4 at failure point

The measured deflection at mid-span versus applied load is plotted in Figure (7.16). The
ductile failure of slab LWC-4 was in agreement with the expected behaviour of slab with the
applied reinforcement ratio. The ultimate capacity was reached at P=49.52 KN, when crushing
of the top compressive surface occurred. The maximum deflection corresponding to the failure
point was 86.65 mm. The failure of the slab LWC-4 occurred in the pure moment zone,
however, the first failure point with the compressive concrete crushing above a crack was not
exactly in the mid-span of the slab.

295
Chapter 7 – Flexural load-deflection behaviour

60

50

40

Load (KN)
30

20

10 LWC-4

0
0 20 40 60 80 100 120
Deflection (mm)

Figure 7.16: Load-deflection curve of slab LWC-4 at mid-span

7.6.7. Comparison of load-deflection curves in all LWC slabs


Figure (7.17) compares the load-deflection behaviour of four lightweight concrete slabs. All
slabs experience a similar trend of ductile failure and there is no significant difference between
the load-deflection curves of slabs.

The average values of the ultimate load-capacity and the maximum deflection are 46.4 KN
and 96 mm respectively. Slabs LWC-2 and LWC-4 are compatible in the recorded data, while
the recorded data of ultimate strength and the maximum deflection in LWC-3 is different
compared to the other slabs. The load-deflection diagram in slab LWC-1 is approximately the
average of all specimens.

60

50

40
Load (KN)

30

20
LWC-1 LWC-2
10
LWC-3 LWC-4

0
0 20 40 60 80 100 120
Deflection (mm)

Figure 7.17: Comparing the load-deflection curves in four LWC slabs


296
Chapter 7 – Flexural load-deflection behaviour

According to Figure (7.17), the elastic load-deflection curve is seen up to 15% of the
ultimate load capacity point. The yielding point of steel bar is about 50% of the ultimate load-
capacity of slab section in almost all slabs.

7.7. PREVIOUS EXPERIMENTS ON LOAD-DEFLECTION BEHAVIOUR


OF SLABS

Nejadi (2005) [1] and Aslani (2014) [2] conducted experimental programs to investigate the
time-dependent behaviour of Conventional Concrete (CC) and Self-Compacting Concrete
(SCC) respectively. The main objective in their study was to monitor the cracking of flexural
elements over long period of time under the effects of both sustained service load and of creep
and shrinkage. They have investigated the load-deflection relationship by applying an increasing
short-term loading up to failure on the slabs. They also have recorded the instantaneous and
long-term deflection of test specimens under service loads and creep and shrinkage effects.

The current study emphasises the time-dependent deflection of flexural elements. To extend
the test results in a wider range of comparable data of previously conducted experiments, this
study used idem dimensions and support conditions of slabs with a different type of concrete. In
other words, to provide a basis of comparison for the deflection variation of the slabs under
different levels of loading, three types of concrete have been used in the slab members with the
same dimensions, bar arrangements and support conditions. By choosing the same number,
diameter, arrangement and cover of steel bars, it is obvious that the other effective parameters
such as the reinforcement ratio bar diameter and bar orientation has no effect on the test results

7.7.1. Short-term loading of conventional concrete slabs (Nejadi, 2005)


Nejadi (2005) [1] investigated the flexural cracking of beams and slabs in short-term and
long-term periods. Twelve prismatic singly-reinforced concrete specimens including six beams
and 6 slabs were cast and moist-cured for 28 days. All specimens were simply supported on a
3.5 m span and tested to failure to investigate the distribution and extent of primary and
secondary cracking under short-term loading.

All specimens were fabricated in 3800 mm long formwork and were simply supported over a
3500 mm span. Eleven Demec targets were glued on the side face of each specimen at the steel
level to measure the concrete surface strains and two targets were fixed onto the top fibre of the
side face at midspan with HBB-X60 non-shrink adhesive. Also, ten electric resistance strain
gauges were attached to one of the main reinforcement bars. Deflection at mid-span was

297
Chapter 7 – Flexural load-deflection behaviour

measured by LVDT. The measured data from LVDT and strain gauges were retrieved through a
data acquisition system connected to a computer, while the load was supplied by a hydraulic
jack connected to an electrically powered pressure pump.

Figure (7.18) shows the schematic diagram of the test set-up for short-term loading of beams
and slabs.

Figure 7.18: Test arrangement for all the beams and slab in short-term loading [1]

Table (7.3) presents the width (b) and height (h) of the sections, number and diameter of
tensile bars for the beams and slabs. All specimens were simply supported along the 3.5 m span
length. There were no compressive bars and stirrups in the sections.

Deflection at mid-span, crack widths, crack patterns, steel strains within the high moment
region and concrete surface strains at the steel level were recorded at each load increment in the
post-cracking range. The development of a primary crack pattern throughout the test was
monitored also.

Table 7.3: Detail of test specimens in the experiment by Nejadi (2005) [1]

Tensile b¯
¯h Tensile b¯
¯h
Beams L(m) 2
Slabs L(m)
bar (mm ) bar (mm2)
B1-a 2N16 3.5 250¯348 S1-a 2N12 3.5 400¯161
B1-b 2N16 3.5 250¯348 S1-b 2N12 3.5 400¯161
B2-a 2N16 3.5 250¯333 S2-a 3N12 3.5 400¯161
B2-b 2N16 3.5 250¯333 S2-b 3N12 3.5 400¯161
B3-a 3N16 3.5 250¯333 S3-a 4N12 3.5 400¯161
B3-b 3N16 3.5 250¯333 S3-b 4N12 3.5 400¯161

Figure (7.19) shows the failure of Beam B1-a and slab S2-b as samples of the conventional
concrete beams and slabs failure in the tests conducted by Nejadi (2005).

298
Chapter 7 – Flexural load-deflection behaviour

(a) (b)
Figure 7.19: Failure of a) slab S2-b and b) beam B1-a [1]

The failure loads and the corresponding deflection for beams and slabs are given in Table
(7.8). Accordingly, the results for slabs are graphically presented in Figure (7.38). The data for
slab S1-B is not available to include in the graph presented in Figure (7.38).

Table 7.4: Summary of the results from short-term flexural test [1]

Failure load Deflection at failure load Failure load Deflection at failure


Slabs
Beams (KN) (mm) (KN) load (mm)
B1-a 109 66 S1-a 22 210
B1-b 103 72 S1-b 21 211
B2-a 112 68 S2-a 37 178
B2-b 104 77 S2-b 36 204
B3-a 138 10 S3-a 50 136
B3-b 127 9 S3-b 47 156

60

50

40
Load (KN)

30

20
S1-A S2-A S2-B

10
S3-A S3-B

0
0 50 100 150 200 250
Deflection (mm)

Figure 7.20: Short-term load-defection curves for conventional concrete slabs

299
Chapter 7 – Flexural load-deflection behaviour

Except for the ratio of tensile reinforcement in the section, all slabs are similar. However,
according to Figure (7.20), there are considerable differences in the load-deflection trend, the
failure load, and the corresponding deflection to the failure load. In slabs S3-a and S3-b with the
highest ratio of tensile reinforcement, the failure load and the corresponding deflection values
are the highest and lowest values of all slabs respectively.

The reinforcement ratio in slabs S2 and S3 are 50% and 100% higher than the reinforcement
in slabs S1, however, the average failure load in slabs S2 and S3 are 69% and 125% higher than
that of slab S1.

7.7.2. Short-term loading of self-compacting concrete slabs (Aslani, 2014)


The experiment of Nejadi (2005) [1] for short-term and long-term monitoring of cracking in
conventional concrete beam and slab was followed by Aslani (2014) [2] with similar specimens
of Self-Compacting Concrete (SCC) slabs.

In the experimental investigation by Aslani (2014) [2], eight singly-reinforced SCC and
Fibre Reinforced Self-Compacting Concrete (FRSCC) slab specimens were cast and moist-
cured for 28 days. All the specimens were simply supported on a 3.5 m span and tested to
failure to investigate the distribution and extent of primary and secondary cracking under short-
term loading. The objectives are similar to the objectives of Nejadi (2005) [1] and the same
dimension slabs were investigated under short-term loading.

Deflections at mid-span, crack widths, crack patterns, steel strains within the high moment
region, and concrete surface strains at the steel level were recorded at each load increment in the
post-cracking range and development of the primary crack pattern was monitored throughout
the test duration. Reinforcement and cover details and the SCC slab dimensions are explained in
Table (7.9).

Table 7.5: Detail of test specimens in the experiment by Aslani (2014) [2]

Slabs Tens. bar Cover(mm) L(m) ¯h (mm2)



NSCC-A 4N16 19 3.5 400¯161
N-SCC-B 4N16 19 3.5 400¯161
D-SCC-A 4N16 19 3.5 400¯161
D-SCC-B 4N16 19 3.5 400¯161
S-SCC-A 4N16 19 3.5 400¯161
S-SCC-B 4N16 19 3.5 400¯161
DS-SCC-A 4N16 19 3.5 400¯161
DS-SCC-B 4N16 19 3.5 400¯161

300
Chapter 7 – Flexural load-deflection behaviour

Figure (7.21) shows a general view of test set-up for short-term loading of SCC and FRSCC
slabs. The instrumentations process and measurement method of the strain and midspan
deflection are similar to those in the experiment by Nejadi (2005) [1].

Figure 7.21: General view of the loading cells, concrete strain gauges and LVDT set-up [2]

Table (7.6) shows a summary of the load-deflection result of SCC slabs under short-term
loading. Figure (7.22) shows the load-deflection curves of the slab at mid-span. Except for the
inclusion or type of the applied fibre in the concrete mixture, all SCC and FRSCC slabs are
similar. In general, the failure load in the FRSCC slab is higher than the normal SCC slab series.
However, the average value of deflection at failure point is high in the FRSCC slabs.

Table 7.6: Summary of the results from short-term flexural test

NSCC- N-SCC- D-SCC- D-SCC- S-SCC- S-SCC- DS-SCC- DS-SCC-


Slab
A B A B A B A B
Failure load (KN) 49 48.5 53 52 50 48 56 54
Deflection at
180 163 205 177 185 167 220 182
failure load (mm)

In Table (7.6), N-SCC, D-SCC, S-SCC and DS-SCC refer to SCC mixtures with no fibre,
steel fibre, polypropylene, and hybrid fibre, respectively.

301
Chapter 7 – Flexural load-deflection behaviour

60

50

Load (KN) 40

30
D-SCC-A D-SCC-B

20 DS-SCC-A DS-SCC-B

N-SCC-A N-SCC-B
10
S-SCC-A S-SCC-B

0
0 50 100 150 200 250
Deflection (mm)

Figure 7.22: Load-deflection curve for slab SCC and FRSCC slabs at mid-span

7.8. COMPARISON OF LOAD-DEFLECTION CURVES IN


LIGHTWEIGHT, CONVENTIOANL AND SELF-COMAPCTING
CONCRETE SLABS

Figure (7.23) compares the load-deflection curves in LWC, SCC and CC slabs. The figure
includes all slabs with different mix design and various ratios of reinforcement.

60

50

40
Load (KN)

30
D-SCC-A D-SCC-B
DS-SCC-A DS-SCC-B
20 N-SCC-A N-SCC-B
S-SCC-A S-SCC-B
LWC-1 LWC-2
LWC-3 LWC-4
10 S1-A S2-A
S2-B S3-A
S3-B
0
0 50 100 150 200 250
Deflection (mm)

Figure 7.23: Comparison of the load-deflection curves in all LWC, SCC and CC slabs

302
Chapter 7 – Flexural load-deflection behaviour

To have a better comparison, the SCC slabs with fibre reinforcing and the CC slabs with
2N12 and 3N12 reinforcement are eliminated from the diagrams in Figure (7.23). Figure (7.24)
shows the comparison between four LWC slabs (LWC, A, B, C, D), two SCC slabs (N-SCC A,
B) and two CC slabs (S3-A, B) in which the slab dimensions and steel bar details are the same.

60

50

40
Load (KN)

30

20 N-SCC-A N-SCC-B
LWC-1 LWC-2
10 LWC-3 LWC-4
S3-A S3-B

0
0 50 100 150 200
Deflection (mm)

Figure 7.24: Load-deflection curves in identical LWC, SCC and CC slabs

Figure (7.25) compares the average load-deflection curves of the LWC, SCC and CC slabs.
The failure load in all slabs are in the same range, however, the failure trend is completely
different in LWC slabs. The failure in CC and SCC slabs are extensively ductile. The LWC
slabs fail with less ductility, compared to the ductile failure of CC and SCC slabs.

The elastic range of the load-deflection curve in all slabs is very close. In other words, they
experience a similar trend of load-deflection up in the elastic range.

303
Chapter 7 – Flexural load-deflection behaviour

60

50

40
Load (KN)
30

20
SCC

10 LWC
CC
0
0 50 100 150 200
Deflection (mm)
Figure 7.25: Comparison of the load-deflection curves in LWC, SCC and CC slabs

According to Figures (7.9), (7.11), (7.13), (7.15), (7.19) and (7.21), there is noticeable
difference in cracking pattern of the slabs under increasing load in terms of the crack spacing
and crack width. The average crack spacing in SCC and CC slabs are less than in EPS-LWC
slabs. However, the crack width and height in the SCC and CC slabs are higher than those of the
EPS-LWC slabs.

7.9. SUMMARY

The second phase of the experimental program, i.e. short-term flexural deflection of the slabs
is discussed in this chapter. The experimental program for short-term load-deflection tests are
explained in detail. The results were compared with the corresponding previously conducted
experiments by Nejadi (2005) on the conventional concrete slabs and Aslani (2014) on the self-
compacting concrete slabs.

Testing setup and slab details

The total number of four one-way lightweight concrete slabs was cast and moist-cured for 28
days. The slab specimens were each nominally 3500 mm long by 400 mm wide and 161 mm
deep. Each slab was reinforced with 4N12 steel bars in tensile side of the section and the
concrete cover was 19 mm. Then each slab was slowly and gradually loaded to failure over a
period of approximately four hours.

304
Chapter 7 – Flexural load-deflection behaviour

All specimens were tested to failure to investigate the mid-span deflection, changes of strain
in critical bending region and distribution and extent of the primary and secondary cracking
under short-term loading using two equal point loads applied at the third points of the span, at
ages greater than 28 days. To make an accurate estimation, the ultimate load capacity of the
section was calculated before the test. The load was then applied by 3 KN increments in each
step on the slab specimens. This rate of loading was kept constant until about 60% of the
expected ultimate load capacity. Then, the load was gradually increased monotonically to failure
and crack patterns were recorded at each load step.

Load-deflection behaviour of lightweight concrete slabs

Table (7.7) describes the ultimate failure load and the maximum deflection of lightweight
concrete slabs under short-term loading. Since all slabs were similar, the average values of the
failure load and maximum deflection at midspan are given also.

Table 7.7: Failure load and maximum deflection of lightweight concrete slabs

Slab Ultimate load (KN) Maximum deflection (mm)

Slab LWC-1 46.5 95.9

Slab LWC-2 48 88

Slab LWC-3 41.53 114.7

Slab LWC-4 49.52 86.65

Average 46.4 96.3

Figure (7.26) compares the load-deflection diagrams in the lightweight concrete slabs.
Variation of the deflection at mid-span by increasing the applied load is similar in all slabs.
Slabs LWC-2 and LWC-4 are compatible in the recorded data, while the recorded data of
ultimate strength and the maximum deflection in LWC-3 is different compared to the other
slabs. The load-deflection diagram in slab LWC-1 is approximately the average of all
specimens.

According to Figure (7.26), the elastic load-deflection curve is seen up to 15% of the
ultimate load capacity point. The yielding point of steel bar is about50% of the ultimate load-
capacity of the slab section in almost all slabs.

305
Chapter 7 – Flexural load-deflection behaviour

60

50

40

Load (KN) 30

20
LWC-1 LWC-2
10
LWC-3 LWC-4
0
0 20 40 60 80 100 120
Deflection (mm)

Figure 10.26: Comparing the load-deflection curves in four LWC slabs

Comparing the results with the previous experiments

Figure (10.27) compares the average load-deflection diagram of the light-weight concrete
slab of this study with the previously performed tests on 8 self-compacting concrete slabs by
Aslani (2014) and 2 conventional nonconcrete slabs by Nejadi (2005). All slabs with three types
of concrete were similar in terms of the geometry and support conditions.

Table (7.8) shows a summary of the load-deflection result of all slabs in three different
experiments under short-term loading.

Table 7.8: Summary of the results from short-term flexural test

Failure Deflection at
Failure Deflection at
Slabs load Failure load
Beams load (KN) failure load (mm)
(KN) (mm)

NSCC-A 49 180 LWC-1 46.5 95.9


concrete slabs
Self-compacting concrete slabs

Lightweight

NSCC-B 48.5 163 LWC-2 48 88

D-SCC-A 53 205 LWC-3 41.53 114.7

D-SCC-B 52 177 LWC-4 49.52 86.65


S-SCC-A 50 185
concrete slabs
Conventional

S-SCC-B 48 167 S3-a 50 136


DS-SCC-A 56 220 S3-b 47 156
DS-SCC-B 54 182

306
Chapter 7 – Flexural load-deflection behaviour

60

50

40
Load (KN)
30

20
SCC

10 LWC
CC
0
0 50 100 150 200
Deflection (mm)
Figure 7.27: Comparison of the load-deflection curves in EPS-LWC, SCC and CC slabs

The failure load in all slabs is in the same range; however, the failure trend is completely
different in LWC slabs. The failure in the CC and SCC slabs are extensively ductile, while,
LWC slabs fail with less ductility, compared to the ductile failure of CC and SCC slabs. The
elastic range of the load-deflection curve in all slabs is very close. In other words, they
experience a similar trend of load-deflection up in the elastic range.

Considering the ratio of tensile reinforcement in the section, all slabs experienced ductile
failure under the increasing flexural loading. In other words, the ultimate bending capacity of
the slab section was more sensitive to the yield strength of the steel bars rather than the
compressive strength of the concrete. According to Table (7.8), there is no considerable
difference in the failure loads of SCC, CC and EPS-LWC slabs and all failure loads are in the
range of 50 KN. While

Despite similar range of the failure load in all slabs, there is a substantial difference in the
maximum deflection of slabs at mid-span at failure load.

Fibre reinforcing of the SCC mixture improves the ductility of the slabs. The ultimate
flexural deflection of these slabs is considerably higher than the other slabs.

Compared to SCC and CC slabs, the lower modulus of elasticity in EPS-LWC slabs resulted
in smaller amount of the maximum deflection at failure load. The relatively brittle failure of the
EPS-LWC slabs may be considered as the reason of the difference in the recorded ultimate
deflection. It emphasises the simultaneous effect of failure load and maximum deflection in
design of one-way slabs.

307
Chapter 7 – Flexural load-deflection behaviour

REFERENCES

1. Nejadi, S., Time-dependent cracking and crack control in reinforced concrete structures. 2005,
PhD thesis, The University of New South Wales Sydney, Australia.
2. Aslani, F., Experimental and numerical study of time-dependent behaviour of reinforced self-
compacting concrete slabs. PhD dissertation, University of Technology Sydney (UTS), 2014.

308
Chapter 8 – Flexural instantaneous deflection

CHAPTER 8

EXPERIMENTAL PROGRAM (PHASE II) –


FLEXURAL INSTANTANEOUS DEFLECTION
Chapter 8 – Flexural instantaneous deflection

Table of Contents

8.1. INTRODUCTION ...................................................................................................................... 309


8.2. EXPERIMENTAL PROGRAM ................................................................................................. 311
8.2.1. Construction of the specimens and curing method ............................................................. 311
8.2.2. Load configuration and arrangement.................................................................................. 312
8.2.3. Monitoring points and measurement .................................................................................. 312
8.2.4. Ultimate capacity and the applied load .............................................................................. 314
8.3. EXPERIMENTAL INSTANTANEOUS DEFLECTION .......................................................... 315
8.4. DIFFERENCE BETWEEN THE ELASTIC AND INSTANTANEOUS DEFLECTION ......... 319
8.5. EFFECT OF LOADING LEVEL ............................................................................................... 321
8.6. PREVIOUS EXPERIMENTS ON THE INSTANTANEOUS DEFLKECTION OF SLABS ... 323
8.6.1. Conventional Concrete Slabs by Nejadi (2005) ................................................................. 323
8.6.2. Self-Compacting Concrete Slabs by Aslani (2014) ............................................................ 324
8.7. COMPARISON OF LOAD-DEFLECTION DIAGRAMS IN LIGHTWEIGHT, SELF-
COMPACTING AND CONVENTIONAL CONCRETE SLABS .......................................................... 325
8.7.1. Instantaneous and Elastic Deflection.................................................................................. 327
8.7.2. Effect of Loading Level on the Flexural Deflection ........................................................... 328
8.8. PROPOSED ANALYTICAL MODEL ...................................................................................... 330
8.9. PROPOSED MODEL OF Ie IN SELF-COMPACTING CONCRETE SLABS ......................... 334
8.10. ROPOSED MODEL OF Ie IN LIGHTWEIGHT CONCRETE SLABS ................................ 335
8.11. SUMMARY ........................................................................................................................... 336
REFERENCES......................................................................................................................................... 341
Chapter 8 – Flexural instantaneous deflection

List of Figures
Figure 8.1: Bar arrangement, cover details and dimension of slab specimens ......................................... 313
Figure 8.2: Loading blocks and LVDT positioning for flexural test of LWC slabs ................................. 313
Figure 8.3 : Monitoring points of flexural instantaneous deflection in LWC slabs ................................. 314
Figure 8.4: Development of compressive strength and modulus of elasticity with time .......................... 315
Figure 8.5 : Placement order of the loading blocks on slabs .................................................................... 316
Figure 8.6: Gradually increment of deflection under each loading block and the instantaneous deflection
of slab LWC-1 after applying the loading blocks .................................................................................... 316
Figure 8.7: Gradually increment of deflection under each loading block and the instantaneous deflection
of slab LWC-2 after applying the loading blocks .................................................................................... 317
Figure 8.8: Gradually increment of deflection under each loading block and the instantaneous deflection
of slab LWC-3 after applying the loading blocks .................................................................................... 318
Figure 8.9: Gradually increment of deflection under each loading block and the instantaneous deflection
of slab LWC-4 after applying the loading blocks .................................................................................... 318
Figure 8.10: Ratio of the recorded deflection to the estimated elastic deflection at 14 and 28 days ........ 320
Figure 8.11: Schematic diagram of instantaneous and elastic deflection of LWC slabs in bending ........ 321
Figure 8.12: (Δins / Δe) 14 under different ratios of service moment and ultimate moment ..................... 321
Figure 8.13: Growing rate of Δins-14 due to 10 % increment of Ma/Mu .................................................... 322
Figure 8.14: Effect of moment ratios on (Δins / Δe) 14 ............................................................................... 323
Figure 8.15: Loading of convention concrete slabs for time-dependent deflection monitoring (Nejadi,
2005) ........................................................................................................................................................ 323
Figure 8.16: General view of the test specimens of SCC and FRSCC slabs (Aslani, 2013) .................... 325
Figure 8.17: Compressive strength and modulus of elasticity of mixtures at 14 and 28 days .................. 327
Figure 8.18: Increasing rate of compressive strength and modulus of elasticity of mixtures................... 327
Figure 8.19: Ratio of recorded instantaneous deflation to elastic deflection at the ages of 14 and 28 days
................................................................................................................................................................. 328
Figure 8.20: Instantaneous / elastic deflection ratio at 14 days under different Ma /Mu ratios ............... 329
Figure 8.21: Effect of moment ratios on Δins / Δe at 14 days .................................................................... 329
Figure 8.22: Effect of moment ratios on instantaneous deflection of slabs at 14 days............................. 330
Figure 8.23: Comparing the experimental and calculated data of instantaneous deflection of slabs ....... 332
Figure 7.24: Comparing the experimental and predicted I e of slabs ........................................................ 333
Figure 8.25: Comparison of the recorded and predicted deflection of slabs at the age of 14 days .......... 335
Figure 8.26: Comparison of the recorded and predicted deflection of slabs at 14 days by proposed model
................................................................................................................................................................. 336
Figure 8.27: Effect of moment ratios on the ratio of instantaneous to elastic deflection of slabs at 14 days.
................................................................................................................................................................. 337
Chapter 8 – Flexural instantaneous deflection

List of Tables
Table 8.1: Loading values, moment ratios and elastic deflection of slabs ............................................... 315
Table 8.2: Instantaneous and elastic deflection of LWC slabs at monitoring points ................................ 320
Table 8.3: Experimental instantaneous deflection of CC slabs (Nejadi, 2005) ........................................ 324
Table 8.4: Experimental instantaneous deflection of SCC and FRSCC slabs (Aslani, 2014) .................. 325
Table 8.5: Loading values, recorded and elastic deflection ..................................................................... 326
Table 8.6: Existing models of effective moment of inertia in RC section ............................................... 331
Table 8.7: Coefficients and limitations of proposed equation for Ie estimation ....................................... 334
Table 8.8: Coefficients and limitations of proposed model for Ie in LWC slabs ..................................... 335
Table 8.9: Instantaneous and elastic deflection of LWC slabs at monitoring points ................................ 337
Table 8.10: Loading values, measured instantaneous and predicted elastic deflection of all slabs .......... 338
Table 8.11: Coefficients and limitations of proposed model for I e in LWC slabs .................................... 339
Table 8.12: Coefficients and limitations of proposed equation for I e estimation ..................................... 339
Chapter 8 – Flexural instantaneous deflection

Chapter 8: Flexural instantaneous deflection

8.1. INTRODUCTION

Performance of reinforced concrete structures under sustained loads particularly in the post-
cracking range is a key design concern. If sections are designed according to strength
requirements only, the behaviour of the structure under sustained service loads may be
unsatisfactory. For example, at service loads, deflections of the member may be excessively
large or the width of the cracks may be improper; nevertheless the degree of the safety against
collapse is acceptable.

For a concrete structure to be serviceable; cracking must be controlled and deflection must
not be excessive. Service load behaviour depends primarily on the properties of the concrete
and these are often not known reliably at the design stage, so it is necessary to check the
serviceability of the structure in the post-cracking range. The behaviour in this range is
complicated by the effects of several factors, which are difficult to assess from analytical
considerations only. Of prime importance are the effects of tension stiffening, random
development of primary and secondary cracks, and the degree of bond breakdown [1].

According to the existing literature, the increase in crack width occurs at a decreasing rate
with time due to long-term or cyclic loading and this increase can be up to twice the initial value
within a few years.

Instantaneous deflection (Δins) is the fundamental parameter of the time-dependent and total
deflection calculation of the Reinforced Concrete (RC) structures. The existing models predict
the Δins for Conventional Concrete (CC) beams considering the concrete properties at the age of

309
Chapter 8 – Flexural instantaneous deflection

28 days [2] and do not cover other types of concrete and loading ages that are inevitable in most
construction projects.

In recent decades, utilizing the chemical and mineral admixtures in concrete technology has
introduced Light-Weight Concrete (LWC) as a reliable and economic construction material.
Using LWC may result in the lighter elements and consequently, the smaller dimension that
both decrease the total weight of the structure and the lateral load effects [3, 4]. Therefore, the
construction cost can be protected when applied to structures such as long span bridge and high-
rise buildings [5, 6]. Also, the better steel-concrete bond, better thermal insulation, durability
performance, tensile strain capacity, and fatigue resistance make it superior to normal weight
concrete [4, 5].

Generally, two types of natural (pumice, diatomite, volcanic cinders, etc.) and artificial
(perlite, clay, sintered fly ash, expanded shale, etc.) lightweight aggregates are used to supply
LWC [7]. Access to lightweight aggregates such as expanded clay, shale, and slate is limited in
some locations and Expanded Polystyrene (EPS) is commercially available worldwide [8]. EPS
is a type of artificial lightweight aggregate with a very low density. It is a thermoplastic foam
involving a gas phase in a polymer matrix and has high compressibility, therefore, it may
provide little restraint to volume changes of the cement paste due to the applied load as well as
the changes in the moisture content [9].

Deflection of flexural elements with low reinforcement ratios is highly sensitive to the age of
concrete at loading and the shrinkage-restraint stress [10, 11]. Modulus of Elasticity (MoE),
loading magnitude and load distribution and support conditions are the principle factors in Δins
of one-way slabs. Also, continuity of the sections along the span; i.e. variable cross-sections and
hence variable moment of inertia depending on the cracking pattern, strongly influence Δins.

Considering the effect of mechanical properties on deflection behaviour of the RC structures,


application of the existing models for LWC should be handled with care. Partial replacement of
normal-weight fine or coarse aggregate with EPS aggregate decreases the Compressive Strength
(CS), MoE and density of LWC depending on the level of replacement [9]. However, the CS in
this type of concrete is more sensitive to density changes than the variation of MoE [7-9].

Mechanical properties of concrete change from an approximately liquid state to a visco-


plastic material within a few hours, which is followed by further development into a hardened
material with almost elastic properties up to about 40% of the compressive strength value. The
cracking moment that has a considerable effect on time-dependent deflection can be reduced by
shrinkage caused during the curing period before 28 days. The loading of young concretes
structures require using the distinct properties at the age of loading; in other words, either the
model or the early age properties should be modified to get an accurate prediction. In LWC
310
Chapter 8 – Flexural instantaneous deflection

containing EPS aggregate, the CS and MoE at seven days is about 75% to 83% and 85% of
those for 28 days, respectively [7-9, 12].

Flexural stiffness in the instantaneous deflection investigation changes with the cracking due
to shrinkage strains of concrete, and the loading magnitude changes the predictable shrinkage
values. The higher the EPS aggregate content in the mix, the greater the drying shrinkage and
cracking development [9], [12].

The main objectives of this Chapter are:

(a) To investigate the measured experimental values of instantaneous deflection of LWC


slabs at the age of 14 days (Δins-14). The LWC contains EPS beads as light-weight aggregate in
the mix;

(b) To compare the development of MoE and CS with time in LWC mix containing EPS
aggregate;

(c) To evaluate and compare the recorded deflections with the predictions of the existing
models of effective moment of inertia (Ie) in codes of practice and empirical equations in the
literature;

(d) To investigate the effect of loading level, ratio of applied service moment (Ma) to
cracking moment (Mcr) and ultimate bending capacity of section (Mu) on Δins-14;

(e) To determine the best matching models of Ie to propose a new model for LWC.

8.2. EXPERIMENTAL PROGRAM

8.2.1. Construction of the specimens and curing method

Despite the detailed and comprehensive explanation of the testing program, construction of
the slabs and companion specimens, curing and loading conditions in Chapter Seven, a brief
summary of the testing program to monitor and record the instantaneous deflection of LWC
slabs is given in this chapter.

The experimental program consists of four LWC one-way slabs to record the deflection and
companion test specimens to measure the mechanical properties. The slabs were simply
supported with identical span length and section dimensions (3.5m length ¯ 0.4m width ¯
0.161m thickness). Each slab specimen was moist cured for 14 days and then subjected to
loading. Six 150mm ¯ 300mm cylindrical specimens to measure CS and three 150mm ¯
300mm cylindrical specimens were used to measure MoE at the different ages. The specimens
for CS and MoE tests were kept covered in a controlled chamber at 20 ± 2oC for 24 hours until
311
Chapter 8 – Flexural instantaneous deflection

demoulding. The relative humidity of the environment was maintained at 50% during the tests.
Then, the specimens were placed in water pre-saturated with lime at 20oC. The compressive
strength test was performed according to AS 1012.14 (1991) [13] and ASTM C39 (2000) [14]
instructions, and the loading rate of the machine on the cylinders was 0.3 MPa/s until failure.
MoE tests confirmed the instructions in AS 1012.17 (2014) [15] and ASTM C469 (2000) [16].

8.2.2. Load configuration and arrangement

In total, four LWC one-way slabs with idem mix design in the concrete laboratory of
University of Technology Sydney (UTS) were investigated. All slabs were internally reinforced
with 4 N12 deformed steel bars. As shown in Figure (8.1), the slabs are similarly simply
supported having length 3500 mm, depth 161 mm and width 400 mm. The side and bottom
cover of concrete from the bar centroid are 40 and 25 mm respectively.

Each pair of slab specimens were subjected to different gravity loads, consisting of self-
weight plus superimposed sustained loads via carefully constructed and arranged concrete
blocks supported on the top surface of the specimens. The uniformly distributed loads are about
30 and 40 percent of the ultimate bending capacity of section in each pair of slabs constructed
with the same mix design of LWC.

To provide the required loading, rectangular concrete blocks of predetermined size and
weights were cast and weighed before the commencement of the test. Slab specimens were
uniformly loaded by the concrete blocks using wooden timbers as loading pads. The wooden
pads, having length 450 mm, depth 30 mm and width 100 mm were placed in equal intervals to
provide uniformly distributed loading on the slabs. Figure (8.2) shows the arrangement of
concrete blocks on the top surface of each slab to achieve the desired sustained load level and
the position of LVDTs under the slabs.

8.2.3. Monitoring points and measurement

The instantaneous deflection of slabs is recorded in two different points along the span at the
age of 14 days. As shown in Figure (8.2), the deflection at mid-span as the maximum flexural
deflection was recorded by Linear Variable Differential Transformer (LVDT). Another LVDT
installed on the border of the high-moment region of the slab recorded the instantaneous
deflection of the slab at the boundary of critical bending moment area along the span length. In
this chapter, Δ1 and Δ2 denote the flexural deflection of slab at mid-span and the border of high-
moment region respectively. Comparing the recorded deflection at these two points provides an

312
Chapter 8 – Flexural instantaneous deflection

overview of the changes in the curvature of slab under the uniformly distributed flexural
deflection. It can also give an insight into the variation of section stiffness along the span.

cs=40 mm, cb=25 mm, s=106 mm 4N12 bars in the


section

Figure 8.1: Bar arrangement, cover details and dimension of slab specimens

Figure 8.2: Loading blocks and LVDT positioning for flexural test of LWC slabs

Figure (8.3) shows the monitoring points along the span length of the slab.

313
Chapter 8 – Flexural instantaneous deflection

Figure 8.3 : Monitoring points of flexural instantaneous deflection in LWC slabs

8.2.4. Ultimate capacity and the applied load

The instantaneous deflection was recorded by LVDT installed at mid-span (Δ1), immediately
after loading of the slabs at the age of 14 days. The ultimate bending capacity (Mu) of a
reinforced concrete section is calculated based on the mechanical properties of concrete at the
age of 28 days. However, the slabs in this study are subjected to the service load for
instantaneous and long-term deflection investigations at the age of 14 days. Mechanical
properties of concrete such as CS and MoE are developing by increasing the age of concrete;
especially up to age of 28 days. Therefore, calculation of the ultimate bending capacity, Mu at
loading age (14 days) by properties of the concrete at the age of 28 days will bring
overestimated capacity and consequently, apply some design errors in deflection calculations.

The existing models give under or overestimated values of the mechanical properties of
concrete. Besides, the models of predicting the properties of LWC at the ages other than 28 days
are very rare in the literature. Therefore, exact values of CS and MoE of the LWC mix have
been measured by test specimens in laboratory conditions at two ages of 14 and 28 days. Figure
(8.4) compares the measured values and development of CS and MoE with time. There are
about 11% and 9% increment of CS and MoE of LWC with time, respectively.

The graph in Figure (8.4) shows the development of CS and MoE between 14 and 28 days
age of concrete only. These graphs are part of nonlinear development trend comprehensively
explained in Chapter Five of this study.

314
Chapter 8 – Flexural instantaneous deflection

35
Compressive strength (MPa)
Modulus of elasticity (GPa)
30

Mechanical properties
25

20

15
10 15 20 25 30
Test age (Day)

Figure 8.4: Development of compressive strength and modulus of elasticity with time

Table (8.1) presents the applied service moment (Ma) of slabs at mid-span due to two
different load conditions of service loads (wa). Also, to consider the exact capacity of slab
section at loading age, the Mu of slab section is presented by applying the mechanical properties
of concrete at two ages of 14 and 28 days. The applied load in each pair of slabs is about 30%
and 40% of the ultimate bending capacity of slabs. In other words, two levels of uniformly
distributed loads are applied to the slabs. The loading levels equal to 30% and 40% of the Mu is
presented by LL1 and LL2 in this Chapter, respectively.

Table 8.1: Loading values, moment ratios and elastic deflection of slabs

wa Ma ࢌᇱࢉି૛ૡ Mu (kN.m) Ma/ Mu (%)


Slab
kN/m kN.m (MPa) 14 days 28 days 14 days 28 days
LWC-A 7.0 10.72 30.96 27.4 27.65 39.13 38.78
LWC-B 7.0 10.72 30.96 27.4 27.65 39.13 38.78
LWC-C 5.29 8.095 30.96 27.4 27.65 29.55 29.29
LWC-D 5.29 8.095 30.96 27.4 27.65 29.55 29.28

8.3. EXPERIMENTAL INSTANTANEOUS DEFLECTION

To attain the required loading levels, concrete blocks of predetermined size and weight were
put on top of the slabs in equal distances. Details of these loading blocks are explained in
Chapter 7 of this study. Figure (8.5) shows the placement order of the concrete blocks in all
slabs.

315
Chapter 8 – Flexural instantaneous deflection

Figure 8.5 : Placement order of the loading blocks on slabs

Figures (8.6) to (8.9) show the instantaneous deflection of slabs under the applied loads.
Since the loading was applied by four concrete blocks on the top surface of the concrete slabs,
there is a sharp increment of deflection after putting each block on the slabs. After all loading
blacks were applied and the changes in the deflection were not considerable, it was considered
as the instantaneous deflection point.

Figure (8.6) shows the gradually increment of the deflection of Slab LWC-1 until reaching
the establishment point that is considered as the instantaneous deflection at two monitoring
points. According to Table (8.1), the applied loading level in this slab is about 0.4Mu (LL2).

14

12

10
Deflection (mm)

8
Δ1
6
Δ2
4

0
0 50 100 150 200 250 300 350 400
Time (min.)

Figure 8.6: Gradually increment of deflection under each loading block and the instantaneous deflection of
slab LWC-1 after applying the loading blocks

According to Figure (8.6), the loading block have similar effect on both monitoring points to
suddenly increase the deflection before establishment point. After the last block placement, the
instantaneous deflection at mid-span grows more than the other point.

316
Chapter 8 – Flexural instantaneous deflection

Figure (8.7) shows the gradually increment of the deflection of Slab LWC-2 until reaching
the establishment point that is considered as the instantaneous deflection at two monitoring
points. According to Table (8.1), the applied loading level in this slab is nearly 0.4Mu (LL2).

14

12

Deflection (mm) 10

6
Δ1
4 Δ2

0
0 50 100 150 200 250 300 350 400
Time (min.)
Figure 8.7: Gradually increment of deflection under each loading block and the instantaneous deflection of
slab LWC-2 after applying the loading blocks

In Slab LWC-2, the effect of loading blocks on the sharp increment of the flexural deflection
is changing considerably after putting the third block on top of the slab.

Figure (8.8) shows the gradually increment of the deflection of Slab LWC-3 until reaching
the establishment point that is considered as the instantaneous deflection at two monitoring
points. According to Table (8.1), the applied loading level in this slab is about 0.3Mu (LL1).

317
Chapter 8 – Flexural instantaneous deflection

Deflection (mm)
5

4 Δ1
Δ2
3

0
0 50 100 150 200 250 300 350 400
Time (min.)

Figure 8.8: Gradually increment of deflection under each loading block and the instantaneous deflection of
slab LWC-3 after applying the loading blocks

Despite different weight of the loading blocks in LL1 in this slab, the increment of the
deflection at two monitoring points is similar to that of slab LWC-2 under LL2.

Figure (8.9) shows the gradually increment of the deflection of Slab LWC-4 until reaching
the establishment point that is considered as the instantaneous deflection at two monitoring
points. According to Table (8.1), the applied loading level in this slab is about 0.3Mu (LL1).

8
7
6
Deflection (mm)

5
4
3
2 Δ1
1 Δ2

0
0 50 100 150 200 250 300 350 400
Time (min.)

Figure 8.9: Gradually increment of deflection under each loading block and the instantaneous deflection of
slab LWC-4 after applying the loading blocks

318
Chapter 8 – Flexural instantaneous deflection

Considering the same loading level on the slabs LWC-3 and LWC-4, the initial deflection
growth and the instantaneous deflection after establishment of changes is similar in these slabs.

8.4. DIFFERENCE BETWEEN THE ELASTIC AND INSTANTANEOUS


DEFLECTION

Considering the instantaneous deflection (Δins) as the elastic deflection (Δe) is a source of
errors in the design of concrete structures, this study compares the experimental data with the
calculated elastic deflection (Δe) of slabs by 14 and 28 days and MoE at two loading levels.

Figure (8.10) shows the ratio of measured instantaneous deflection to the predicted elastic
deflection at the ages of 14 days (Δins-14 /Δe-14) and the ratio of recorded instantaneous to the
elastic deflection at the age of 28 days (Δins-14 /Δe-28) in the LWC slabs. The Δe is calculated on
the basis of the gross moment of inertia in the uncracked section at mid-span.

Equations (8.1) and (8.2) are utilized to determine the elastic deflection of slabs under two
levels of loading values and Mcr respectively.

ͷ ܹ௔ Ǥ ݈ ସ
߂௘ ൌ ൈ ‫ݍܧ‬Ǥ ሺͺǤͳሻ
͵ͺͶ ‫ܧ‬௖ Ǥ ‫ܫ‬௚

‫ܯ‬௖௥ ൌ ‫ܫ‬௖௥ ൈ ݂௥ ‫ݍܧ‬Ǥ ሺͺǤʹሻ

Where; Wa, l, Ec, Ig, Icr and fr are uniformly distributed load (KN/m), span length (m), MoE of
concrete (KN/m2), gross moment of inertia (m4), cracking moment of inertia (m4) and modulus
of rupture (KN/m2), respectively.

319
Chapter 8 – Flexural instantaneous deflection

3
Δ-inst14/Δ-e14
2.5
Δ-inst14/Δ-e28
2

Δ-inst/Δ-e
1.5

0.5

0
LWC-1 LWC-2 LWC-3 LWC-4
LWC slabs
Figure 8.10: Ratio of the recorded deflection to the estimated elastic deflection at 14 and 28 days

The ratio of (Δins / Δe) 14 in the mid-span of slabs varies between 1.58 and 2.39. Comparing
the recorded deflection at 14 days (Δins-14) with the estimated elastic deflection at 28 days (Δe-28),
the ratio of Δins-14 / Δe-28 shows a discrepancy between 1.73 and 2.62. In fact, the increased
flexural stiffness at the age of 28 days decreases the elastic deflection and, consequently,
enhances the ratio of Δins-14 / Δe-28. The average ratio of (Δins / Δe) 14 and Δins-14 / Δe-28 are 1.98 and
2.17 respectively.

Table (8.2) shows the measured instantaneous and calculated elastic deflection of LWC slabs
at the two monitoring points of slab. The Δ1 and Δ2 are the measured instantaneous deflections
respectively at two monitoring points, while Δe1 and Δe2 are the calculated elastic deflection of
slabs at the corresponding points.

Table 8.2: Instantaneous and elastic deflection of LWC slabs at monitoring points

Loading level Loading level Δe-1 Δe-2

Slab 0.3Mu 0.4Mu (mm) (mm)


Δ1(mm) Δ2 (mm) Δ1(mm) Δ2 (mm) 14days 28 day 14days 28days

LWC-1 11.84 11.19 4.99 4.57 4.37 4.11

LWC-2 10.76 9.22 4.99 4.57 4.37 4.11

LWC-3 6.33 5.43 3.77 3.45 3.22 2.85

LWC-4 5.98 5.16 3.77 3.45 3.22 2.85

320
Chapter 8 – Flexural instantaneous deflection

Figure (8.11) shows a schematic diagram to compare the calculated elastic and measured
instantaneous deflection of slabs under uniformly distributed loading.

Figure 8.11: Schematic diagram of instantaneous and elastic deflection of LWC slabs in bending

8.5. EFFECT OF LOADING LEVEL

Since the slabs are identical concerning the dimensions, reinforcement and support
conditions, the relative ratio of the applied load to the flexural capacity of section (Ma/Mu) could
be a critical factor in Δins of slabs that are generally lightly reinforced concrete structures. This
ratio becomes more important for the loading of the slabs before 28 days age. Figure (8.12)
compares the effect of Ma/Mu on the ratio of (Δins / Δe) 14 in slabs. The applied service load is
about 30 and 40 percent of the ultimate section capacity in each pair of slabs with the same mix
design.

3
LWC-4
(Δ-inst / Δ-e) at 14 days

LWC-1
2.7
LWC-2
LWC-3
2.4

2.1

1.8

1.5
0 20 40 60
Ration of moments- Ma / Mu (%)

Figure 8.12: (Δins / Δe) 14 under different ratios of service moment and ultimate moment

321
Chapter 8 – Flexural instantaneous deflection

The increasing rate of Δins-14 under two levels of Ma/Mu ratio is different from the variation of
(Δins/Δe)14 ratio. Unexpectedly, the effect of amplifying Ma/Mu ratio by 10% on the Δins-14 is
considerably different in these identical slabs. Increasing the Ma/Mu ratio by 10% has resulted
about 80% to 100% higher Δins-14 in the LWC slabs. In other words, the early-state deflection of
LWC slabs (before the age of 28 days) is significantly sensitive to the loading levels and
variation. Figure (8.13) compares the Δins-14 of slabs under two levels of the Ma/Mu ratio.

14

12

10
Δ-inst-14 (mm)

0
LWC-3 LWC-4 LWC-1 LWC-2
Slab specimens
Figure 8.13: Growing rate of Δins-14 due to 10 % increment of Ma/Mu

Together with the effects of Ma and Mu, cracking moment (Mcr) can be another critical
parameter in deflection calculations of slabs. Notwithstanding the effects of shrinkage, M cr
causes the first cracking in the tensile zone and changes the section behavior to non-elastic.
Since Mcr and Mu are dependent characteristics of the reinforced concrete section, they may
have similar influences on the Δins/Δe ratio. Besides, the ratio of Ma/Mu is limited to 30 to 40% in
each pair of slabs in this study. In fact, it looks like the application of the predefined
relationship between Ma and Mu in this study. According to Figure (8.14) the (Δins/Δe)14 ratio is
changing by variation of Mcr/Mu and Ma/Mu ratios; however, it is noticeably more sensitive to
the ratio of Mcr/Ma compared to the Ma/Mu ratio. It is evident from Figure (8.12) that the Ma/Mu
and Mcr/Ma ratios have increasing and decreasing effects on (Δins /Δe)14 respectively.

The value of the applied Ma is very close to Mcr in LWC-3 and LWC-4, in which the elastic
behavior of the section is expected. However due to the shrinkage effects and initiation of
cracking in the tension zone, Δins is about 60% higher than the estimated elastic deflection.

322
Chapter 8 – Flexural instantaneous deflection

3
Mcr/Ma Mcr/Mu

Moments and deflection ratio


2.5 Ma/Mu Δ-inst/Δ-e

1.5

0.5

0
LWC-1 LWC-2 LWC-3 LWC-4
Slab specimens
Figure 8.14: Effect of moment ratios on (Δins / Δe) 14

8.6. PREVIOUS EXPERIMENTS ON THE INSTANTANEOUS


DEFLKECTION OF SLABS

8.6.1. Conventional Concrete Slabs by Nejadi (2005)

Similar to the short-term program, a total of twelve specimens including six beams and six
slabs, with the same cross section and details for short-term tests, were monitored for up to 400
days. One batch of commercially pre-mixed concrete was used to manufacture all the test and
companion specimens. The steel stain within high moment regions, the concrete surface strains
at the tensile steel level, deflection at the mid-span, crack widths and crack spacing, were
recorded throughout the testing period. All beam and slabs were subjected to different gravity
sustained loads at the age of 14 days. For both the beams and slab specimens, two identical
specimens were constructed for each combination of variables, with one loaded to 40% of its
ultimate capacity (type a) and one loaded to 30% of its ultimate capacity (type b). Figure (8.15)
shows the general view of test arrangements for time-dependent deflection monitoring.

Figure 8.15: Loading of convention concrete slabs for time-dependent deflection monitoring (Nejadi, 2005)

323
Chapter 8 – Flexural instantaneous deflection

The instantaneous deflection of conventional concrete slabs at mid-span was recorded


immediately after the loading blocks were applied on the top surface of slabs. Table (8.3) shows
the instantaneous deflection of conventional concrete slabs under two different levels of service
loading.

Table 8.3: Experimental instantaneous deflection of CC slabs (Nejadi, 2005)

Instantaneous deflection
Slab
Load level 0.26 Mu Load level 0.2 Mu
CC-a 11.8

CC-b 5.04

8.6.2. Self-Compacting Concrete Slabs by Aslani (2014)

A total of eight numbers of Self-Compacting Concrete (SCC) and Fibre Reinforced Self-
Compacting Concrete (FRSCC) slabs with the same cross-section and details as for the short-
term tests were monitored for up to 240 days to measure the time-dependent development of
cracking and deformations under service loads. Deflections at mid-span, crack widths, crack
patterns, steel strains within the high moment region, and concrete surface strains at the steel
level were recorded under the service load and shrinkage effects. For this purpose, four SCC
mixes, two plain SCC, two steel, two polypropylene, and two hybrid FRSCC slab specimens
were considered in the test program. The SCC mixture without fibre, with steel fibre and
polypropylene fibre are called N-SCC, D-SCC and S-SCC respectively. The term DS-SCC
refers to the SCC mixture with both the steel and polypropylene fibres.

Two sustained load levels were considered, namely 40% and 30% of the ultimate design
load, and designated load conditions ‘a’ and ‘b’, respectively. The applied load consisting of
self-weight plus superimposed sustained loads at the age of 14 days via carefully constructed
and arranged concrete blocks supported off the top (of the specimens). Figure (8.16) shows the
general view of test arrangement for time-dependent deflection monitoring.

324
Chapter 8 – Flexural instantaneous deflection

Figure 8.16: General view of the test specimens of SCC and FRSCC slabs (Aslani, 2013)

The instantaneous deflection of SCC and FRSCC slabs at mid-span was recorded
immediately after the loading blocks were applied on the top surface of the slabs. Table (8.4)
shows the instantaneous deflection of SCC and FRSCC slabs under two different levels of
service loading.

Table 8.4: Experimental instantaneous deflection of SCC and FRSCC slabs (Aslani, 2014)

Instantaneous deflection
Slab
Load 0.4 Mu Load 0.3 Mu
N-SCC-a 12.1

N-SCC-b 5.89

D-SCC-a 7.65

D-SCC-b 7.59

S-SCC-a 6.41

S-SCC-b 2.91

DS-SCC-a 8.98

DS-SCC-b 5.14

8.7. COMPARISON OF LOAD-DEFLECTION DIAGRAMS IN


LIGHTWEIGHT, SELF-COMPACTING AND CONVENTIONAL
CONCRETE SLABS

Table (8.5) shows the values of the experimental instantaneous deflection and the calculated
elastic deflection of LWC, SCC and CC slabs under two levels of sustained loading. The values
of Mu at 14 and 28 days, Ma, and the ratio of Ma /Mu are also given in the table.

325
Chapter 8 – Flexural instantaneous deflection

Table 8.5: Loading values, recorded and elastic deflection

Δ ins
wa Mu (KN.m) Ma Ma / Mu (%) Δ e (mm)
Slab
(KN/m) 14days 28days (KN.m) 14 day 28 day 14days 28days (mm)
N-SCC-a 7.31 27.47 27.78 11.189 40.73 40.27 3.18 2.9 12.1

N-SCC-b 5.41 27.47 27.78 8.28 30.14 29.8 2.36 2.15 5.89

D-SCC-a 7.26 27.84 28.05 11.12 39.94 39.65 3.5 2.85 7.65

D-SCC-b 5.36 27.84 28.05 8.21 29.48 29.27 2.58 2.11 7.59

S-SCC-a 7.3 27.73 28.05 11.18 40.32 39.85 3.45 2.87 6.41

S-SCC-b 5.4 27.73 28.05 8.27 29.83 29.48 2.55 2.12 2.91

DS-SCC-a 7.33 28.05 28.34 11.23 40.03 39.63 3.29 2.85 8.98

DS-SCC-b 5.43 28.05 28.34 8.32 29.66 29.36 2.44 2.11 5.14

LWC-1 7 27.4 27.65 10.72 39.13 38.78 4.99 4.57 11.96

LWC-2 7 27.4 27.65 10.72 39.13 38.78 4.99 4.57 11.4

LWC-3 5.29 27.4 27.65 8.095 29.55 29.28 3.77 3.45 6.3

LWC-4 5.29 27.4 27.65 8.095 29.55 29.28 3.77 3.45 5.98

CC-a 4.44 26.03 27.05 6.8 26.12 25.14 2.73 2.5 11.8

CC-b 3.46 26.03 27.05 5.3 20.36 19.59 2.13 1.95 5.04

Figures (8.17) and (8.18) compare the measured values and increasing rate of the
Compressive Strength (CS) and Modulus of Elasticity (MoE) of the mixtures in LWC, SCC and
CC slabs at the ages of 14 and 28 days.

326
Chapter 8 – Flexural instantaneous deflection

50 50
14 days 28 days
14 days 28 days
45 45

Modulus of Elasticity (GPa)


Compressive strength (MPa)
40 40

35 35

30 30

25 25

20 20

15 15
N-SCC D-SCC S-SCC DS-SCC LWC CC N-SCC D-SCC S-SCC DS-SCC LWC CC
Mixtures Mixtures

Figure 8.17: Compressive strength and modulus of elasticity of mixtures at 14 and 28 days

50
N-SCC S-SCC D-SCC 50
N-SCC D-SCC S-SCC
DS-SCC LWC CC
45 45 DS-SCC LWC CC
Compressive strength (MPa)

40 Modulus of elasticity (GPa) 40


35 35
30 30
25 25
20
20
15
10 15 20 25 30
15
Test age
10 15 20 25 30
Test ages

Figure 8.18: Increasing rate of compressive strength and modulus of elasticity of mixtures

8.7.1. Instantaneous and Elastic Deflection

According to Figures (8.17) and (8.18), the average increasing ratio of MoE in CC, SCC and
LWC is 1.10, 1.17 and 1.10 respectively. The increasing ratio of compressive strength in NC,
SCC and LWC is 1.35, 1.15 and 1.10 respectively. Overall the early age properties and
increasing rate in SCC mixtures particularly in presence of fibres are better than the
corresponding properties of CC and LWC.

Figure (8.19) shows the ratio of the experimental instantaneous deflection (Δins) to the elastic
deflection (Δe) at the ages of 14 and 28 days in concrete slabs with different mixture designs.
The ratio of (Δins/Δe)14 in CC, LWC and SCC slabs vary between 1.15 in S-SCC-b to 4.3 in NC-
a. However the majority of ratios are in the range of 1.5-2.5. Increased elasticity modulus at the
age of 28 days decreases the elastic deflection at the age of 28 days and hence increases the
ratio of Δins-14/Δe-28.

327
Chapter 8 – Flexural instantaneous deflection

5
Exp14/Elas-14 Exp14/Elas-28
4.5
4
3.5

Δ-inst/Δ-e 3
2.5
2
1.5
1
0.5
0

Slab Specimn
Figure 8.19: Ratio of recorded instantaneous deflation to elastic deflection at the ages of 14 and 28 days

8.7.2. Effect of Loading Level on the Flexural Deflection

All slab specimens are identical with the same reinforcement arrangement. Among all
effective parameters in instantaneous deflection of lightly reinforced concrete structures, the
loading rate and Ma/Mu ratio play major role. Figure (8.20) compares the effect of Ma/ Mu on the
ratio of (Δins/ Δe)14 in the slabs. The CC slabs are under 20% and 26% of their ultimate capacity,
while SCC and LWC slabs are under about 30% and 40% of the ultimate bending capacity
loading. Except D-SCC (a, b) slabs, a 10% load increase in the slabs, causes about 50 % to 70%
growth in the (Δins/ Δe)14 ratio. Considering the uncertainties and unexpected loads during the
construction, this ratio of (Δins/ Δe)14 becomes more crucial in the accurate design of concrete
slabs. The low strength CC slabs are more sensitive to the loading values, while a 6% load
increment causes about an 80% rise in the (Δins/ Δe)14 ratio.

328
Chapter 8 – Flexural instantaneous deflection

5
4.5 CC-a

Δ (inst) / Δ (Elastic) at 14 days


4 N-SCC-a
3.5
D-SCC-b
3 DS-SCC-a
2.5 N-SCC-b LWC-1

2 CC-b DS-SCC-b NC-a LWC-2 D-SCC-a


LWC-3
1.5 LWC-4 s-scc-a
1 S-SCC-b
0.5
0
10 20 30 40 50
Ms / Mu (%)

Figure 8.20: Instantaneous / elastic deflection ratio at 14 days under different Ma /Mu ratios

Along with the magnitude of the service moment, the effect of cracking moment (Mcr) that
causes the first cracking is considerable in deflection calculations of slabs. Since Mcr and Mu are
dependent characteristics of section, consequently they may change the Δins / Δe ratio in a similar
manner. The ratio of Ma/Mu is limited to 20% to 26% in CC slabs and 30% to 40% in LWC and
SCC slabs; in other words Ma is related to Mu in this experiment. According to Figure (8.21),
the ratio of (Δins/ Δe)14 changes very similarly in various ratios of Mcr/ Mu and Ma/ Mu, however
the ratio of Mcr/ Mu is more effective in instantaneous deflection calculations especially in low
strength concrete specimens.

5
Mcr/Ma Mcr/Mu
Moments and deflection ratio

4.5
4 Ma/Mu Δ-inst/Δ-e
3.5
3
2.5
2
1.5
1
0.5
0
N-SCC-a

D-SCC-a

NC-a

NC-b
DS-SCC-a

DS-SCC-b

LWC-1

LWC-2

LWC-3

LWC-4
N-SCC-b

D-SCC-b

S-SCC-a

S-SCC-b

Slab specimens
Figure 8.21: Effect of moment ratios on Δins / Δe at 14 days

329
Chapter 8 – Flexural instantaneous deflection

In all slabs the applied Ma is smaller than Mcr and the section is supposed to be in elastic
range, however due to the shrinkage-induced cracking, the measured instantaneous deflection is
higher than calculated elastic deflection. As presented in Figure (8.22), the instantaneous
deflection in one way slabs slightly increases by decreasing the ratio of Mcr/Ma, while it
increases by increasing the ratios of Mcr/Mu and Ma/Mu in general.

2.5
Mcr/Ma Mcr/Mu Ma/Mu

2
Moment ratio

1.5
R² = 0.1387
1
R² = 0.0003
0.5 R² = 0.2597

0
2 4 6 8 10 12 14
Instantaneous deflection
Figure 8.22: Effect of moment ratios on instantaneous deflection of slabs at 14 days

8.8. PROPOSED ANALYTICAL MODEL

Existing models of the effective moment of inertia (Ie) have been utilized to predict the Δins-14
of LWC, SCC and CC slabs. Among the existing models, some equations are developed for
particular conditions like point loading and the concrete members with FRP bars and near
surface mounted FRP strips; however, their predictions have been compared with Δins-14 values
in this study to observe any compatibility in between.

Estimated values of Ie from Equations (8.1) to (8.20) (see Table (8.6)) have been utilized to
predict the Δins-14 of slabs. The recorded Δins-14 in slabs have been compared and evaluated with
these predicted deflections in Figure (8.24). Surprisingly, there are considerable differences
between the predicted and experimental data in two levels of the applied Ma / Mu ratio. Existing
models give a better prediction at the higher loading level (Ma/Mu = 0.4), and overestimate the
Δins-14 by about 100% in the lower loading level (Ma/Mu = 0.3).

The effective moment of inertia in the measured instantaneous deflection of slabs is


calculated and compared with the existing models. Considering the identical conditions and
known parameters such as support condition, loading magnitude and mechanical properties of
the section in each slab, the actual value of Ie relevant to the recorded deflection has been
compared with the Ie determined by Equations (8.3) to (8.21) in Figures (8.23) and (8.24).
330
Chapter 8 – Flexural instantaneous deflection

Table 8.6: Existing models of effective moment of inertia in RC section

Reference Equation Equation


ୡ୰ 
ୣ ൌ ሺ ൗ ሻଶ ୥ ൅ ቀͳ െ ሺ ୡ୰ൗ ሻଶ ቁ ୡ୰ ൑ ୥
ୟ ୟ Eq. (8.3)
Bronson (1963) ୡ୰ 
ୣ ൌ ሺ ൗ ሻଷ ୥ ൅ ቀͳ െ ሺ ୡ୰ൗ ሻଷ ቁ ୡ୰ ൑ ୥ Eq. (8.4)
Bronson (1965) [17] ୟ ୟ

ୡ୰  Eq. (8.5)


ୣ ൌ ሺ ൗ ሻସ ୥ ൅ ቀͳ െ ሺ ୡ୰ൗ ሻସ ቁ ୡ୰ ൑ ୥
ୟ ୟ

ୡ୰
Al Shaik and Al Zaid (1993) [18] ୣ ൌ ୡ୰ ൅ ൫ ୥ െ ୡ୰ ൯ሺ ൗ ሻଷି଴Ǥ଼஡ ൑ ୥ Eq. (8.6)

ACI-318-08 [19]
ACI-435-68 [20]
ୡ୰ 
ୣ ൌ ሺ ൗ ሻଷ ୥ ൅ ቀͳ െ ሺ ୡ୰ൗ ሻଷ ቁ ୡ୰ ൑ ୥ Eq. (8.7)
ୟ ୟ
AS-3600-09 [21]

୑ౙ౨ ଷ ୑ౙ౨ ଷ
ୣ ൌ ቀ ቁ Ⱦୢ ୥ ൅ ൤ͳ െ ቀ ቁ ൨ ୡ୰ ൑ ୥
୑౗ ୑౗
ACI 440 (2006) [22] Eq. (8.8)
୉ ஡
Ⱦୢ ൌ Ƚୠ ሺ ౜ ൅ ͳሻ, Ⱦୢ ൌ ͲǤʹ ൏ͳ
୉౩ ஡ౘ

୍ౙ౨
Bischoff and Paixao (2004) [23] ୣ୤ ൌ ୑ మ ൑  ୥
ଵି஗ቀ ౙ౨ൗ୑ ቁ
౗ Eq. (8.9)
Eurocode 2, (1994) [24]
η = 1 - Icr/ Ig
୍ౙ౨
Bischoff (2005, 2007) [25] ୣ ൌ ୑ ൑  ୥
ଵି஗ஒቀ ౙ౨ൗ୑ ቁ
౗ Eq. (8.10)
Bischoff and Scanlon (2007) [25]
no tension stiffening Ͳ ൑ Ⱦ ൑ ͳ full tension stiffening
୍ౙ౨
ୣᇱ ൌ ୑ ൑  ୥
ଵି஗ஓஒቀ ౙ౨ൗ୑ ቁ
Bischoff and Gross (2011) [26] ౗ Eq. (8.11)
γ=1.72 – 0.72Mcr /Ma , β= Mcr /Ma
ଶଷ୍ౙ౨ ୍౛
Faza and GangaRao (1992) [27] ୫ ൌ Eq. (8.12)
଼୍ౙ౨ ାଵହ୍౛

୑ౙ౨ ଷ ୍ౝ ୑ౙ౨ ଷ
Benmokrane et al. (1996) [28] ୣ ൌ ቀ ቁ ൅ ͲǤͺͶ ൤ͳ െ ቀ ቁ ൨ ୡ୰ ൑ ୥ Eq. (8.13)
୑౗ ଻ ୑౗

୑ౙ౨ ଷ ୑ౙ౨ ଷ
ୣ ൌ ቀ ቁ Ⱦୢ ୥ ൅ ൤ͳ െ ቀ ቁ ൨ ୡ୰ ൑ ୥
୑౗ ୑౗
Yost et al. (2003) [29] and
Eq. (8.13)
ACI-440 (2006) [22] ɏ
Ⱦୢ ൌ ͲǤʹ ൏ͳ
ɏୠ

୑ౙ౨ ଷ ୑ౙ౨ ଷ ୍ౙ౨


ୣ ൌ ቀ ቁ Ⱦୢ ୥ ൅ ൤ͳ െ ቀ ቁ ൨ ൑ ୥
୑౗ ୑౗ ஓ

Rafi and Nadjai (2009) [30] ͲǤͲͲͳ͹ɏ ୤ Eq. (8.14)


ɀൌሺ ൅ ͲǤͺͷͶͳሻሺ ൅ ͳሻ
ɏୠ ʹୱ

୑ౙ౨
ୣ ൌ ୡ୰ ˆ‘” ൐͵ Eq. (8.15)
୑౗
Alsayed et al. (2000) [31] ʹ ୡ୰ ୡ୰
ୣ ൌ ൤ͳǤͶ െ ൬ ൰൨ ୡ୰ ൑ ୥ ˆ‘”ͳ ൏ ൏͵
ͳͷ ୟ ୟ Eq. (8.16)
୍ౝ ୍ౙ౨
ISIS Canada (2001) [32] ୣ ൌ ౉ మ ൑  ୥ Eq. (8.17)
୍ౙ౨ ାሾଵି଴Ǥହቀ ౉ౙ౨ ቁ ሿ൫୍ౝି୍ౙ౨ ൯

୍ౝ ୍ౙ౨
Hall and Ghali (2000) [33] ୣ ൌ ౉ మ ൑  ୥ Eq. (8.18)
୍ౙ౨ ାሾଵିஒభ ஒమ ቀ ౉ౙ౨ ቁ ሿ൫୍ౙ౨ ି୍ౝ ൯

୔ౙ౨ ଷ ୡ୰ൗ ଷ
ୣ ൌ ቀ ቁ ୥ ൅ ቀͳ െ ሺ ሻ ቁ ୡ୰ ൑ ୥
ACI- 224.2R-92 [34] ୔ Eq. (8.19)
Ag: Gross cross section area , Acr: nAs
୑ ୐
ିሺ ౗ൗ୑ ሻሺ ౙ౨ൗ୐ሻ஡
ୣ ൌ ୡ୰ ൅ ൫ ୥ െ ୡ୰ ൯ ౙ౨ ൑ ୥ for ρ>1% Eq. (8.20)
Fikry and Thomas (1998) [35]
୑ ୐ Eq. (8.21)
ିሺ ౗ൗ୑ ሻሺ ౙ౨ൗ୐ሻ
 ୣ ൌ ୡ୰ ൅ ൫ ୥ െ ୡ୰ ൯ ౙ౨ ൑ ୥

331
Chapter 8 – Flexural instantaneous deflection

5
N-SCC-a N-SCC-b D-SCC-a D-SCC-b S-SCC-a
4.5 S-SCC-b DS-SCC-a DS-SCC-b LWC-1 LWC-2

Ratio of predicted to real Δins


LWC-3 LWC-4 CC-a CC-b
4

3.5

2.5

1.5

0.5

0
Eq.(8.3)

Eq.(8.4)

Eq.(8.5)

Eq.(8.6)

Eq.(8.7)

Eq.(8.8)

Eq.(8.9)

Eq.(8.10

Eq.(12)

Eq.(8.13
Eq.(8.11)

Eq.(8.14)

Eq.(8.15)

Eq.(8.16)

Eq.(8.17)

Eq.(8.18)
Equations in doces and experiments

Figure 8.23: Comparing the experimental and calculated data of instantaneous deflection of slabs

In Figure (8.23), the measured experimental data of instantaneous deflection in the LWC
slabs are compared with the instantaneous deflection of the same slabs by existing Ie
models in Table (8.6).

In Figure (8.24), the effective moment of inertia for the slabs are calculated with the
existing models in Table (8.6) and compared with the effective moment of inertia for the
measured deflection. The effective moment of inertia for the measured deflections in the
slabs is calculated form the Equation (8.22).

‫ݓ‬௔ Ǥ ݈ ସ
ୣ ൌ Ǥ ‫ݍܧ‬Ǥ ሺͺǤʹʹሻ
‫ܧ‬௖ Ǥ ο௜௡௦

Where; k is the coefficient of support condition, wa is the applied loading, l is the span
length and Δins is the measured instantaneous deflection of slab.

332
Chapter 8 – Flexural instantaneous deflection

2 N-SCC-a N-SCC-b D-SCC-a D-SCC-b S-SCC-a


1.8 S-SCC-b DS-SCC-a DS-SCC-b LWC-1 LWC-2
LWC-3 LWC-4 CC-a CC-b
1.6
1.4
1.2
1
Ie / I real

0.8
0.6
0.4
0.2
0
Eq. (8.3)

Eq. (8.4)

Eq. (8.5)

Eq. (8.6)

Eq. (8.7)

Eq. (8.8)

Eq. (8.9)

Eq. (8.10)

Eq. (8.11)

Eq. (8.12)

Eq. (8.13)

Eq. (8.14)

Eq. (8.15)

Eq. (8.16)

Eq. (8.17)

Eq. (8.18)
Equations in codes and experiments

Figure 7.24: Comparing the experimental and predicted Ie of slabs

Comparative investigation of the presented models and experimental data show that the
models in Equations (8.13) and (8.14) give conservative Δins-14 predictions in SCC and FRSCC
slabs. In this study, certain required intrinsic and/or extrinsic variables (i.e., properties of
concrete, support conditions, loading type and age) in Ie prediction are well covered by the
presented equations. Beside, both Equation (8.3) as a basis of all other related studies and
Equation (8.18) give the closest prediction to the recorded experimental data. Therefore, with
the models in Equations (8.3) and (8.16) as the basis, the effect of time-dependent elasticity,
fibre volume fraction in the mixture, ultimate and cracking moment capacity along with the
applied service moment have been included in the equation in order to obtain a better estimate
of the instantaneous deflection.

Fibre type, fibre volume fraction and the aspect ratio of fibber are the main effective
parameters in comparison of fibre reinforced sections especially in the tensile and flexural
behaviour. Regarding the negligible difference in aspect ratio of the applied fibres in the slabs,
the effect of type and volume fraction of the fibre on the instantaneous deflection is
investigated. Additionally, according to considerable effect of the time-dependent Mu on the
moment ratios and deflection in Figures (8.20) and (8.21), the Mcr/Mu ratio is also included in
the developed relationship. To include the effect of the loading on the immature concrete
(before 28 days), the ratio of modulus of elasticity of concrete at loading age to that of the 28
days is applied in the equation.
333
Chapter 8 – Flexural instantaneous deflection

Equations (8.3), (8.16) are basically developed for beams. Equation (8.16) predicts Ie in two
distinct ranges of 1<Mcr / Ma < 3 and Mcr/Ma >3 and there is no limit for Mcr / Ma in Equation
(8.3). The ratios of Mcr /Ma and Mcr /Mu for slabs in this study are between 0.87-1.48 and 0.35-
0.43 respectively. Accordingly, the proposed equation is recommended to apply for Mcr / Ma< 3.

8.9. PROPOSED MODEL OF Ie IN SELF-COMPACTING CONCRETE


SLABS

Equation (8.23) is proposed to predict the Ie in SCC and FRSCC slabs. The developed model
is based on the most-compatible results in Figure (8.23) and (8.24). The required coefficients
and the limitations of Equation (8.23) application are illustrated in Table (8.7).

ሺଵି଴Ǥଵ୚౜ ሻ
ୡ୰ ெమ
ୣ ൌ Ƚ ൈ ୡ୰ ቀ ൗ ቁ ൅ ቂሺͳ െ Ⱦሻ ൅ ሺெ ೎ೝሻሺమశഁሻ ቃ ୥  Eq. (8.23)
ୟ ೌ ൈெೠ

Table 8.7: Coefficients and limitations of proposed equation for Ie estimation

Coefficient of fibre Ratio of elasticity Fibre content


Ie , ρ, Ma
type ( α ) β ‫܎܄‬
β=Ec-14/ Ec-28
௘  ൑ ͲǤ͸ ୥ α=0.9 for DS-SCC Fibre volume
Ec-14 and Ec-28
α=1.0 for SCC fraction in the mixture
ߩ ൒ ͲǤͲͲͷ Modulus of elasticity of
ୡ୰ α=1.15 for D-SCC (Kg/m3) presented in
ൗ ൏ ͵ concrete at
ୟ α=1.95 for S-SCC Table 1
14 and 28 days

Figure (8.25) shows the comparison of the recorded data of instantaneous deflection for SCC
and FRSCC slabs with the prediction of the proposed equation by application corresponding
coefficients. Obviously the predicted values in all slabs are close to the recorded values.

334
Chapter 8 – Flexural instantaneous deflection

12

10

Predcited Δ-ins (mm)


8

0
0 2 4 6 8 10 12
Recorded Δ-ins (mm)

Figure 8.25: Comparison of the recorded and predicted deflection of slabs at the age of 14 days

8.10. ROPOSED MODEL OF Ie IN LIGHTWEIGHT CONCRETE


SLABS

Equation (8.24) is proposed to predict Ie in LWC slabs. The developed model is based on the
most-compatible results in Figure (8.23) and (8.24). The required coefficients and limitations in
application of the model are illustrated in Table (8.8). This type of LWC is widely used in
Australia; hence, the maximum Ie is recommended to agree with the limitations of AS-3600-09
[36].

‫ܯ‬௖௥ ଶ ሺ‫ܯ‬௖௥ ሻଵǤ଻


‫ܫ‬௘ ൌ ‫ܫ‬௖௥ ൈ ൬ ൰ ൅ ቈሺͳ െ ߚሻ ൅ ቉ ൈ ‫ܫ‬௚ ‫ݍܧ‬Ǥ ሺͺǤʹͶሻ
‫ܯ‬௔ ሺ‫ܯ‬௔ ൈ ‫ܯ‬௨ ሻ

Table 8.8: Coefficients and limitations of proposed model for Ie in LWC slabs

Coefficient β ρ Mcr/Ma Ie
Limit Ec-14/ Ec-28 ≥ 0.005 ≤3 ≤ 0.6Ig

Where; ρ is the ratio of the tensile reinforcement in the section, and Ec-14 and Ec-28 are the
modulus of elasticity of LWC at 14 and 28 days respectively.

Excluding the area of steel bars from the section, underestimates the Ig by 17%. Therefore,
the transformed section method is utilized to minimize the error in the calculation of Ig of the
uncracked section in the slabs.

335
Chapter 8 – Flexural instantaneous deflection

Figure (8.26) compares the recorded Δins-14 of LWC slabs with the prediction of the proposed
model in Equation (8.24). The proposed model accurately predicts the Ie values; consequently,
the predicted Δins-14 of LWC slabs is in good agreement with the measured experimental data.

Figure 8.26: Comparison of the recorded and predicted deflection of slabs at 14 days by proposed model

8.11. SUMMARY

The instantaneous deflection of lightweight concrete slabs subjected to service load in the
experimental program is described in detail. Effect of the loading level and the ratio of applied
moment to the cracking moment of slab on the recorded instantaneous deflection are also
investigated.

The experimental program consists of four LWC slabs used to record the deflection
immediately after applying the uniformly distributed loading on the slabs. The slabs were
simply supported with identical span length and section dimensions (3.5m length ¯ 0.4m width
¯ 0.161m height). Each slab specimen was moist cured for 14 days and then subjected to
loading. All slabs were internally reinforced with 4 N12 deformed steel bars.

The applied load on the slabs consisted of self-weight plus superimposed sustained loads by
the concrete blocks on top of the slabs. Each pair of slabs was subjected to the total loading
equal to 30% and 40% of the ultimate bending capacity of the slab section.

The instantaneous deflection of slabs was recorded in two different points along the span at
the age of 14 days. The first point was the mid-span of the span and second point was the border
of high moment region in the span length.

The measured instantaneous deflection was compared with the calculated elastic deflection of
slabs at age 14 days. Table (8.9) shows the instantaneous and elastic deflection of the LWC
slabs at the monitoring points in mid-span and the third point of slab.
336
Chapter 8 – Flexural instantaneous deflection

Table 8.9: Instantaneous and elastic deflection of LWC slabs at monitoring points

Loading level (1) Loading level (2) Δe-1 (mm) Δe-2 (mm)
Slab
Δ1(mm) Δ2 (mm) Δ1(mm) Δ2 (mm) 14d 28d 14d 28d

LWC-1 11.84 11.19 4.99 4.57 4.37 4.11

LWC-2 10.76 9.22 4.99 4.57 4.37 4.11

LWC-3 6.33 5.43 3.77 3.45 3.22 2.85

LWC-4 5.98 5.16 3.77 3.45 3.22 2.85

According to Figure (8.27) the (Δins / Δe)14 ratio is changing by variation of the Mcr / Mu and
Ma/Mu ratios. However, the (Δins / Δe)14 ratio are it is noticeably more sensitive to the ratio of
Mcr/Ma compared to Ma/Mu ratio. It is evident from Figure (8.27) that the Ma/Mu and Mcr/Ma
ratios have increasing and decreasing effects on (Δins / Δe)14 respectively.

3
Mcr/Ma Mcr/Mu
Moments and deflection ratio

2.5 Ma/Mu Δ-inst/Δ-e

1.5

0.5

0
LWC-1 LWC-2 LWC-3 LWC-4
Slab specimens
Figure 8.27: Effect of moment ratios on the ratio of instantaneous to elastic deflection of slabs at 14 days.

Comparing the results with previously conducted experiments

The experimental results of instantaneous deflection of lightweight concrete slabs are


analysed and discussed broadly and compared with the previously conducted experimental
results of conventional concrete slabs by Nejadi (2005) and self-compacting concrete slabs by
Aslani (2014).

Table (8.10) shows the recorded instantaneous deflection in the lightweight, self-compacting
and conventional concrete slabs. The applied service load and service moment and the
calculated elastic deflection at two ages of 14 and 28 days are given in the table.
337
Chapter 8 – Flexural instantaneous deflection

Table 8.10: Load values, measured instantaneous and predicted elastic deflection of slabs

wa Ma Δe (mm) Δins
Slab
(KN/m) (KN.m) 14 day 28 day (mm)
N-SCC-a 7.31 11.189 3.18 2.9 12.1
N-SCC-b 5.41 8.28 2.36 2.15 5.89
D-SCC-a 7.26 11.12 3.5 2.85 7.65
D-SCC-b 5.36 8.21 2.58 2.11 7.59
S-SCC-a 7.3 11.18 3.45 2.87 6.41
S-SCC-b 5.4 8.27 2.55 2.12 2.91
DS-SCC-a 7.33 11.23 3.29 2.85 8.98
DS-SCC-b 5.43 8.32 2.44 2.11 5.14
LWC-1 7 10.72 4.99 4.57 11.96
LWC-2 7 10.72 4.99 4.57 11.4
LWC-3 5.29 8.095 3.77 3.45 6.3
LWC-4 5.29 8.095 3.77 3.45 5.98
CC-a 4.44 6.8 2.73 2.5 11.8
CC-b 3.46 5.3 2.13 1.95 5.04

Comparison of the results presented in Table (8.10) gives the following conclusions:

- Except the D-SCC (a, b) slabs, 10% increase of the service load causes about 100%
increment of the instantaneous deflection in the slabs of all concrete types. In D-SCC (a,
b) slabs, the increased load shows almost no effect on the instantaneous deflection at the
age of 14 days;
- The instantaneous deflection at the first load level (0.3Mu in LWC and SCC slabs, 0.2Mu
in CC slabs) in all slab specimens except S-SCC is in a similar range;
- The instantaneous deflection at the higher load level (0.4Mu in LWC and SCC slabs,
0.26Mu in CC slabs) in N-SCC, LWC and CC slabs is in a similar range, however, the
instantaneous deflection in S-SCC is about half that of the other slabs;
- There is considerable difference between the predicted elastic deflection and measured
instantaneous deflection of all slabs. The ratio of instantaneous to elastic deflection varies
from 1.14 in S-SSC-b to 3.8 in N-SCC-a.

Equations (10.1) and (10.2) are proposed and verified to predict Ie in lightweight and
conventional concrete and self-compacting concrete slabs with and without fibre reinforcement
respectively. The required coefficients and the limitations of the Equations (8.25) and (8.26)
application are illustrated in Tables (9.11) and (9.12) respectively.

‫ܯ‬௖௥ ଶ ሺ‫ܯ‬௖௥ ሻଵǤ଻


‫ܫ‬௘ ൌ ‫ܫ‬௖௥ ൈ ൬ ൰ ൅ ቈሺͳ െ ߚሻ ൅ ቉ ൈ ‫ܫ‬௚ ‫ݍܧ‬Ǥ ሺͺǤʹͷሻ
‫ܯ‬௔ ሺ‫ܯ‬௔ ൈ ‫ܯ‬௨ ሻ

338
Chapter 8 – Flexural instantaneous deflection

Table 8.11: Coefficients and limitations of proposed model for Ie in LWC slabs

Coefficient β ρ Mcr/Ma Ie
Limit Ec-14/ Ec-28 ≥ 0.005 ≤3 ≤ 0.6Ig

ሺଵି଴Ǥଵ୚౜ ሻ
ୡ୰ ெమ
ୣ ൌ Ƚ ൈ ୡ୰ ቀ ൗ ቁ ൅ ቂሺͳ െ Ⱦሻ ൅ ሺெ ೎ೝሻሺమశഁሻ ቃ ୥ “Ǥ ሺͺǤʹ͸ሻ
ୟ ೌ ൈெೠ

Table 8.12: Coefficients and limitations of proposed equation for Ie estimation

Coefficient of fiber Ratio of elasticity Fiber content


Ie , ρ, Ma
type ( α ) β ‫܎܄‬
β=Ec-14/ Ec-28
௘  ൑ ͲǤ͸ ୥ α=0.9 for DS-SCC Fiber volume
Ec-14 and Ec-28
ߩ ൒ ͲǤͲͲͷ α=1.0 for SCC fraction in the mixture
Modulus of elasticity of
ୡ୰ α=1.15 for D-SCC (Kg/m3) presented in
ൗ ൏ ͵ concrete at
ୟ α=1.95 for S-SCC Table 1
14 and 28 days

According to the instantaneous deflection of three different types of reinforced concrete slabs
made with normal, lightweight and self-compacting concrete slabs, the following conclusion
can be drawn from the current study:

- Flexural deflection behavior of slabs is strongly influenced by concrete type with similar
compressive strength;
- In the existing models in the literature, there is an almost linear relationship between the
load magnitude and the instantaneous deflection of slabs. The results of this study
indicated a considerable nonlinear effect of the load level on the instantaneous deflection,
especially in the low-strength concrete slabs;
- 10% increment of the load value causes about 100% and 105% higher instantaneous
deflection in the lightweight concrete and self-compacting concrete slabs, respectively. In
low-strength normal concrete slabs, the instantaneous deflection is raise about 135% by
6% higher loading;
- The instantaneous deflection of fiber reinforced self-compacting concrete slabs
experienced less variations with the changes of loading level;
- Considering the elastic deflection equal to instantaneous deflection even for Ma < Mcr in
slabs is completely incorrect and causes unsafe design of reinforced concrete slabs.
- Elastic deflection is not a trustworthy indicator of the instantaneous, time-dependent and
total deflection of slabs;
- Development of the mechanical properties of normal, lightweight and self-compacting
concrete with age is different before the age of 28 days. Therefore, for the early age

339
Chapter 8 – Flexural instantaneous deflection

loading, calculations of the instantaneous deflection of the slabs based on the mechanical
properties at the age of 28 days may result in inaccurate values;
- The ratio of service moment to the ultimate moment capacity of slabs strongly affects the
instantaneous deflection;
- A new model is proposed and verified to predict the effective moment of inertia in the
self-compacting concrete slabs with and without fiber reinforcing. The predicted
instantaneous deflections by this model are in agreement with the experimental results;
- Another model of the effective moment of inertia is proposed and verified for the normal
and lightweight concrete slabs. The experimentally recorded instantaneous deflections
are well-predicted by this model.

340
Chapter 8 – Flexural instantaneous deflection

REFERENCES

1. Nejadi, S., Time-dependent cracking and crack control in reinforced concrete structures. 2005,
PhD thesis, The University of New South Wales Sydney, Australia.
2. Mazzotti, C. and M. Savoia, Long-term deflection of reinforced self-consolidating concrete
beams. ACI Structural Journal, 2009. 106(06).
3. Yoğurtcu, E. and K. Ramyar, Self-compacting lightweight aggregate concrete: design and
experimental study. Magazine of Concrete Research, 2009. 61(7): p. 519-527.
4. Güneyisi, E., M. Gesoğlu, and E. Booya, Fresh properties of self-compacting cold bonded fly
ash lightweight aggregate concrete with different mineral admixtures. Materials and structures,
2012. 45(12): p. 1849-1859.
5. Mazaheripour, H., et al., The effect of polypropylene fibers on the properties of fresh and
hardened lightweight self-compacting concrete. Construction and Building Materials, 2011.
25(1): p. 351-358.
6. Vakhshouri, B. and S. Nejadi, Mix design of light-weight self-compacting concrete. Case Studies
in Construction Materials, 2016. 4: p. 1-14.
7. Xu, Y., et al., Mechanical properties of expanded polystyrene lightweight aggregate concrete
and brick. Construction and Building Materials, 2012. 27(1): p. 32-38.
8. Trussoni, M., C.D. Hays, and R.F. Zollo, Fracture Properties of Concrete Containing Expanded
Polystyrene Aggregate Replacement. ACI Materials Journal, 2013. 110(5).
9. Sabaa, B. and R.S. Ravindrarajah. Engineering properties of lightweight concrete containing
crushed expanded polystyrene waste. in Materials Research Society, 1997, Fall Meeting,
Symposium MM, Advances in Materials for Cementitious Composites December. 1997.
10. Scanlon, A. and P.H. Bischoff, Shrinkage restraint and loading history effects on deflections of
flexural members. ACI Structural Journal, 2008. 105(4).
11. Vakhshouri, B. and S. Nejadi. Experimental study of time dependent bond transfer length under
pure tension in slabs. in International Conference on Construction Materials and Structures.
2014. IOS Press BV.
12. Li, Y., N. Liu, and B. Chen, Properties of lightweight concrete composed of magnesia phosphate
cement and expanded polystyrene aggregates. Materials and Structures, 2015. 48(1-2): p. 269-
276.
13. AS-1012.14, Methods of testing concrete - Method for securing and testing cores from hardened
concrete for compressive strength. Standards Australia, 1991.
14. ASTM-C39, Standard Test Method for Compressive Strength of Cylindrical Concrete
Specimens. american Society of Testing and Materials, 2000.
15. AS-1012.17-97, Methods of testing concrete - Determination of the static chord modulus of
elasticity and Poisson's ratio of concrete specimens. Standards Australia, 2014.
16. ASTM-C469/C469M-14, Standard Test Method for Static Modulus of Elasticity and Poisson’s
Ratio of Concrete in Compression. American Society of Testing and Materials, 2000.
17. Vakhshouri, B. and S. Nejadi. Limitations and Uncertainties in the Long Term Deflection
Calculation of Concrete Structures. in Second International Conference on Vulnerability and
Risk Analysis and Management (ICVRAM) and the Sixth International Symposium on
Uncertainty, Modeling, and Analysis (ISUMA). 2014.
18. Alshaikh, A.H. and R. Al-Zaid, Effect of reinforcement ratio on the effective moment of inertia
of reinforced concrete beams. ACI Structural Journal, 1993. 90(2).
19. ACI-318-02. Building code requirements for structural concrete (ACI 318-08) and commentary.
2002. American Concrete Institute, International Organization for Standardization.
20. ACI-435, Deflections of reinforced concrete flexural member. American Concrete Institute
(ACI), Concrete International, 1968.

341
Chapter 8 – Flexural instantaneous deflection

21. AS-3600-09, Concrete structures. Australian Standard, 2009.


22. ACI-440-06, Guide for the design and construction of concrete reinforced with FRP bars.
American Concrete Institute (ACI) Committee 440, ACI 440.1R-06, ACI, Farmington Hills, MI,
44, 2006.
23. Bischoff, P.H. and R. Paixao, Tension stiffening and cracking of concrete reinforced with glass
fiber reinforced polymer (GFRP) bars. Canadian Journal of Civil Engineering, 2004. 31(4): p.
579-588.
24. du Béton, C.E.-I., CEB-FIP model code (MC-90). 1993, Thomas Telford Ltd, London.
25. Bischoff, P.H., Deflection calculation of FRP reinforced concrete beams based on modifications
to the existing Branson equation. Journal of Composites for Construction, 2007. 11(1): p. 4-14.
26. Bischoff, P.H. and S.P. Gross, Equivalent moment of inertia based on integration of curvature.
Journal of Composites for Construction, 2010. 15(3): p. 263-273.
27. Faza, S. and H. GangaRao. Pre-and post-cracking deflection behaviour of concrete beams
reinforced with fibre-reinforced plastic rebars. in Proceedings of the First International
Conference on Advance Composite Materials in Bridges and Structures (ACMBS-I), Canadian
Society of Civil Engineers, Sherbrooke, Cananda. 1992.
28. Benmokrane, B., O. Chaallal, and R. Masmoudi, Flexural response of concrete beams
reinforced with FRP reinforcing bars. ACI Structural Journal, 1996. 93(1): p. 46-55.
29. Yost, J.R., S.P. Gross, and D.W. Dinehart, Effective moment of inertia for glass fiber-reinforced
polymer-reinforced concrete beams. ACI Structural Journal, 2003. 100(6).
30. Rafi, M.M., et al., Aspects of behaviour of CFRP reinforced concrete beams in bending.
Construction and Building Materials, 2008. 22(3): p. 277-285.
31. Alsayed, S., Y. Al-Salloum, and T. Almusallam, Performance of glass fiber reinforced plastic
bars as a reinforcing material for concrete structures. Composites Part B: Engineering, 2000.
31(6): p. 555-567.
32. Newhook J., Reinforcing concrete structures with fibre reinforced polymers. ISIS Canada:
Design Manual No.3, The Canadian Network of Centres of excellence on intelligent sensing for
innovative structures, Winnipeg, Manitoba, Canada, 2001.
33. Hall, T. and A. Ghali, Long-term deflection prediction of concrete members reinforced with
glass fibre reinforced polymer bars. Canadian Journal of Civil Engineering, 2000. 27(5): p. 890-
898.
34. ACI-224.2R-86, Cracking of concrete members in direct tension American Concrete Institute
(ACI) Committee 224, Detroit, USA, 1986.
35. Fikry, A.M. and C. Thomas, Development of a model for the effective moment of inertia of one-
way reinforced concrete elements. ACI Structural Journal, 1998. 95(4).
36. AS-3600-09, Concrete structures. Standards Australia, 2009.

342
Chapter 9 – Flexural long-term deflection

CHAPTER 9

EXPERIMENTAL PROGRAM (PHASE II) –


FLEXURAL LONG-TERM DEFLECTION
Chapter 9 – Flexural long-term deflection

Table of Contents

9.1. INTRODUCTION....................................................................................................................... 343


9.2. EXPERIMENTAL PROGRAM ................................................................................................. 346
9.2.1. Construction of the Specimens and Curing ........................................................................ 346
9.2.2. Monitoring points of time-dependent flexural deflection ................................................... 347
9.2.3. Time-dependent Deflection of Lightweight Concrete Slabs .............................................. 348
9.2.4. Testing Duration ................................................................................................................. 349
9.2.5. Experimental Data .............................................................................................................. 349
9.2.6. Time-Dependent Deflection at Monitoring Points ............................................................. 350
9.3. RATIO OF LONG-TERM TO INSTANTANEOUS DEFLECTION ........................................ 355
9.4. RATE OF TIME DEPENDENT DEFLECTION IN LIGHTWEIGHT CONCRETE SLABS ... 357
9.5. CRACKING HISTORY OF THE LIGHTWEIGHT CONCRETE SLABS AT EARLY AGES OF
LOADING................................................................................................................................................ 358
9.6. SIMIALR STUDIES ON THE CONVENTIONAL AND SELF-COMPACTING CONCRETE
SLABS 360
9.6.1. Conventional Concrete Slabs .................................................................................................. 360
9.6.2. Self-Compacting Concrete Slabs ............................................................................................ 361
9.7. COMPARISON OF TIME-DEPENDENT DEFLECTION IN LIGHTWEIGHT, SELF-
COMPACTING AND CONVENTIOANL CONCRETE SLABS .......................................................... 363
9.8. TIME-DEPENDENT DEFLECTION IN CODES OF PRACTICE ........................................... 366
9.9. RATIO OF LONG-TERM TO INSTANTANEOUS DEFLECTION IN LIGHTWEIGHT, SELF-
COMPACTING AND CONVENTIONAL CONCRETE SLABS .......................................................... 368
9.10. DEFLECTION RATE IN LIGHTWEIGHT, SELF-COMPACTING AND CONVENTIONAL
CONCRETE SLABS ............................................................................................................................... 374
9.11. SUMMARY ............................................................................................................................ 375
REFERENCES ......................................................................................................................................... 383
Chapter 9 – Flexural long-term deflection

List of Figures

Figure 9.1: Cross section and reinforcing detail of LWC slab specimens in long-term deflection test .... 346
Figure 9.2: General view of the long-term deflection test set up of LWC slabs ...................................... 347
Figure 9.3: Loading pads and concrete block on the slabs for time-dependent testing ............................ 348
Figure 9.4: Illustrative sustained loads slab specimens ............................................................................ 349
Figure 9.5: Time-dependent deflection of slab LWC-A under LL2 at monitoring points ........................ 350
Figure 9.6: Time-dependent deflection of slab LWC-B under LL2 at monitoring points ......................... 351
Figure 9.7: time-dependent deflection variation of slab LWC-C under LL1 at monitoring points ........... 351
Figure 9.8: Time-dependent deflection variation of slab LWC-D at monitoring points .......................... 352
Figure 9.9: Time-dependent deflection of LWC-A and LWC-B slabs under LL2 and LWC-C and LWC-D
under LL1 at mid-span .............................................................................................................................. 352
Figure 9.10: Time-dependent deflection of LWC-A and LWC-B slabs under LL2 and LWC-C and LWC-
D under LL1 at the border of high moment region ................................................................................... 353
Figure 9.11: Ratio of deflection at mid-span to deflection at border point of high moment region under
LL1 (slabs LWC-C and LWC-D) and LL2 (slabs LWC-A and LWC0B) ................................................. 354
Figure 9.12: Ratios of deflection at corresponding points of LWC-A under LL2 and LWC-C under LL1
.................................................................................................................................................................. 354
Figure 9.13: Ratio of long-term to instantaneous deflection in LWC slabs under LL2 (0.4 Mu ) ............. 356
Figure 9.14: Ratio of long-term to instantaneous deflection in LWC slabs underLL1 (0.3Mu) ................ 356
Figure 9.15: Deflection rate of slabs C and D under LL1 (0.3Mu)............................................................ 358
Figure 9.16: Deflection rate of slabs A and B under LL2 (0.4Mu)............................................................ 358
Figure 9.17: Cracking pattern of slab LWC-A at 7 and 48 days after sustained loading ......................... 359
Figure 9.18: Cracking pattern of slab LWC-B at 7 and 48 days after sustained loading.......................... 359
Figure 9.19: Cracking history of slab C at 7 and 48 days after sustained loading .................................... 359
Figure 9.20: Cracking history of slab D at 7 and 48 days after sustained loading ................................... 359
Figure 9.21: The loading of convention concrete slabs for time-dependent deflection monitoring ......... 361
Figure 9.22: Time-dependent deflection of slabs at mid-span (Nejadi, 2005) ......................................... 361
Figure 9.23: The loading of Self-compacting concrete slabs for time-dependent deflection monitoring
(Aslani, 2014) ........................................................................................................................................... 362
Figure 9.24: Time-dependent deflection of SCC and FRSCC slabs under 0.4Mu loading (Aslani, 2014)
.................................................................................................................................................................. 362
Figure 9.25: Time-dependent deflection of SCC and FRSCC slabs under 0.3Mu loading (Aslani, 2014)
.................................................................................................................................................................. 363
Figure 9.26: Comparison of deflection vs. time in SCC, LWC and CC slabs under higher loading level364
Figure 9.27: Comparison of deflection vs. time in SCC, LWC and CC slabs under lower loading level 365
Figure 9.28: Deflection vs. time in LWC and CC slabs under higher loading level ................................ 365
Figure 9.29: Predicted Δins and Δtot of SCC, LWC and CC slabs under a) LL1, b) LL2 ........................... 367
Figure 9.30: The ratio of long-term to instantaneous deflection under higher loading level .................... 369
Figure 9.31: The ratio of long-term to instantaneous deflection under lower loading level ..................... 370
Figure 9.32: Comparing the experimental and predicted ratio of the total to instantaneous deflection ... 372
Figure 9.33: The ratio of long-term to instantaneous deflection of SCC and FRSCC slabs (higher loading
level)......................................................................................................................................................... 372
Figure 9.34: The ratio of long-term to instantaneous deflection of SCC and FRSCC slabs (lower loading
level)......................................................................................................................................................... 373
Figure 9.35: The ratio of long-term to instantaneous deflection of all slabs (higher loading level) ......... 373
Figure 9.36: Deflection rate of all slabs under higher loading level ......................................................... 374
Figure 9.37: Deflection rate of all slabs under lower loading level .......................................................... 374
Figure 9.38: Cross section and reinforcing detail of LWC slab specimens in long-term deflection test .. 375
Figure 9.39: Time-dependent deflection of LWC-A and LWC-B slabs under LL2 and LWC-C and LWC-
D under LL1 at mid-span ......................................................................................................................... 376
Figure 9.40: Time-dependent deflection of LWC-A and LWC-B slabs under LL2 and LWC-C and LWC-
D under LL1 at the border of high moment region ................................................................................... 376
Chapter 9 – Flexural long-term deflection

Figure 9.41: Comparison of deflection vs. time in SCC, LWC and CC slabs under higher loading level378
Figure 9.42: Comparison of deflection vs. time in SCC, LWC and CC slabs under lower loading level 379
Figure 9.43 The ratio of long-term to instantaneous deflection under higher loading level ..................... 379
Figure 9.44: The ratio of long-term to instantaneous deflection under lower loading level ..................... 380
Figure 9.45: Comparing the experimental and predicted ratio of the total to instantaneous deflection ... 381
Figure 9.46: Predicted Δins and Δtot of SCC, LWC and CC slabs under a)LL1, b) LL2 ............................ 382
Chapter 9 – Flexural long-term deflection

List of Tables

Table 9.1: loading level and bending capacity of the LWC slabs in the long-term test ........................... 350
Table 9.2: Ratio of total to instantaneous deflection of LWC slabs at different ages .............................. 355
Table 9.3: Loading level, predicted elastic deflection and recorded instantaneous deflection of slabs .... 364
Table 9.4: Calculation of long-term deflection in codes of practice......................................................... 366
Table 9.5: Long-term to short-term deflection ratio in slab series tested by Nejadi and Gilbert (2005) .. 368
Table 9.6: Long-term to short-term deflection ratio in SCC and FRSCC slabs ....................................... 368
Table 9.7: Ratio of long-term to instantaneous deflection in LWC, CC and SCC slabs .......................... 369
Table 9.8: Free shrinkage strain and creep coefficient of CC, LWC and SCC mixtures.......................... 370
Table 9.9: loading level and bending capacity of the LWC slabs in the long-term test ........................... 375
Table 9.10: Ratio of total to instantaneous deflection of LWC slabs at different ages ............................ 377
Table 9.11: Ratio of the long-term to instantaneous deflection of LWC, SCC and CC slabs .................. 378
Table 9.12: Free shrinkage strain and creep coefficient of CC, LWC and SCC mixtures ........................ 380
Chapter 9 – Flexural long-term deflection

Chapter 9: Flexural long-term deflection

9.1. INTRODUCTION

There have been a considerable number of investigations about the time-dependent behaviour
of concrete structures in the literature since it was first experienced and reported almost a
century ago. However, there is no distinctive boundary in Reinforced Concrete (RC) members
as to when the instantaneous (short-term) deformation ends and the long-term deformation
begins. Even, there is no end for time-dependent deflection of reinforced concrete structures
during the service life which in turn, makes the prediction of the final total deflection
inaccurate. To determine the long-term deflection of a concrete member, the time-dependent
effects of shrinkage, creep and the consequence reduction in tension stiffening need to be
accurately considered. Additional factors which can contribute to the increased long-term
deflection include the development of new cracks, widening of secondary cracks due to creep
and shrinkage effects, and effects of the repeated load cycles [1]. The initial deflection in the
flexural reinforced concrete elements can be considerably smaller than the long-term values,
thus, the time-dependent deflection is fundamentally significant in design and assessment of
reinforced concrete structures.

Instantaneous deflection (Δins) is the basis of time-dependent and total deflection calculation
of RC structures. As by definition, long-term deflection is the increase in instantaneous
deflection by time due to inelastic deformations as a result of creep and shrinkage effects. The
ratio of long-term to the instantaneous deflection is a crucial parameter in serviceability
considerations of the reinforced concrete structures. Although it is not verified by all
experimental investigations, the existing design codes predict the long-term deflection of
flexural elements as a multiple of the initial or instantaneous deflection. It might generally lead

343
Chapter 9 – Flexural long-term deflection

to an underestimated time-dependence and consequently, total deflection of the RC member.


Therefore, it may reduce the safety margin in the design process dramatically.

Along with the uncertainties associated with determination of creep, shrinkage and material
properties of the concrete, accurate prediction of Δins is a major problem in the calculation of
long-term and total deflection of slabs [2]. Existing models in ACI-209-08 [3] and ACI-318-08
[4] overestimate the modulus of elasticity and the effective stiffness and therefore,
underestimate the instantaneous and long-term deflection in the concrete slabs subjected to early
age loading [5-7].

Long-term deflections are calculated either by modifying the material parameters such as
modulus of elasticity or solely by multiplying the initial deflection by a factor. Deflection
multipliers depend on the increase in curvature due to cracking, shrinkage and creep which in
turn depend on (a) structural layout; (b) the ratio of applied load to cracking load; (c)
reinforcement; and (d) material properties [8].

Rahman et al. (1998) [9] proposed a simplified method of predicting the long-term deflection
of cracked beams by the multiplication of an amplification factor in the initial deflection. The
multiplier factor was a function of the long-term strain and the initial neutral axis.

In the finite element analysis, some researchers apply the reduced modulus of elasticity to
predict the time-dependent deflection of slabs to account for the deflection increment by time
due to cracking, creep and shrinkage. Vollum (2003) [8] verified the previously suggested
equivalent elastic modulus model for the flat slabs and found a factor of approximately “2” for
the increased long-term deflection due to the effects of cracking and shrinkage.

Neville et al. (1977) [10] discussed the limitation of deflection control by span to depth ratio
in some codes of practice and compared the results with predictions of British Standard
CP110.1972. They also investigated the accuracy of the long-term deflection multiplier in some
design codes and found that the creep and shrinkage do not increase the calculated instantaneous
deflections as much as where multipliers are used. The calculated long-term deflections by
CP110.1972 [11] were of the order of 1.5 to 2 times the instantaneous values and the span to
depth ratio gave underestimated long-term deflection due to the effects of creep and shrinkage.

Deflection of flexural elements with low-reinforcement ratio e.g. slabs is highly sensitive to
the loading before the age of 28 days [12]. Modulus of elasticity, loading value, load
distribution and support conditions are the principle factors in the deflection calculation of slabs.
Also, continuity of sections along the span; i.e. variable cross-sections and hence variable
moment of inertia depending on the degree of cracking, strongly influence the instantaneous and
time-dependent deflections.

344
Chapter 9 – Flexural long-term deflection

Mechanical properties of concrete are changing continuously from an approximately liquid


state to a visco-plastic material within a few hours, which is followed by further development
into a hardened material with almost elastic properties at two loading values up to about 40% of
the compressive strength of concrete. In different types of concrete, the compressive strength
and modulus of elasticity at seven days is changing from about 50% to 85% of those for 28 days
respectively [13-16]. The early age loading requires using the distinct properties at the age of
loading; in other words, either the model or the early age properties should be modified to get an
accurate prediction. The flexural stiffness is also changing with the cracking due to creep and
shrinkage strains of concrete, and the early age loading changes the predictable creep and
shrinkage behaviour.

Gilbert and Guo (2005) [17] performed a three year experimental program by testing of seven
specimens of large-scale reinforced concrete flat slabs. They recorded the ratio of long-term to
instantaneous deflection in the range of 4 to 7 after 240 days and 5 to 9 after 1100 days, while,
this ratio may be as large as two to three in the predictions of design codes [18]. The loading
history and the environmental effects were a source of considerable differences in the slabs.

In the absence of a reliable reference, the unknown effects of early age loading and type of
concrete make the deflection estimation of reinforced concrete slabs more complicated.
Determination and limitation of the time-dependent deflection by Δins play actively a critical role
in RC slab design. Hence accurate estimation of Δins and Δlong enriches the total deflection
estimation and consequently makes an improved structural and performance-based design.

The validity of the constitutive laws for the instantaneous and long-term behaviour of CC,
LWC and SCC slabs should be verified, and the effect of mixture proportions must be carefully
checked. This study also provides empirical evidence of the principal impact of the creep
coefficient and shrinkage strain on the long-term deflection of slabs made by different types of
concrete.

The main objectives of this chapter are:

- Presenting the time-dependent deflection of Light-Weight Concrete (LWC) slabs;

- Comparing the experimental values of instantaneous and time-dependent deflection of


conventional Concrete (CC), Self-Compacting Concrete (SCC) and LWC slabs subjected
to the loading at the age of 14 days (Δins-14, Δlong);

- Investigating the effect of the loading level, creep and shrinkage strains on Δins-14 and the
ratio of Δlong / Δins-14;

345
Chapter 9 – Flexural long-term deflection

- Comparing the experimental results of the instantaneous and long-term deflection with
the predictions of ACI-318-08 [4], AS-3600-09 [19], and CSA.A23.04 [20].

9.2. EXPERIMENTAL PROGRAM

9.2.1. Construction of the Specimens and Curing

The lightweight concrete slabs described in Chapter (7) were maintained under the sustained
loading to monitor the time-dependent deflection at different points. In other words, the slabs
for the instantaneous study were kept under the applied loads to monitor and record the long-
term deflection due to effects of service loading, creep and shrinkage.

Figure (9.1) shows the reinforcing detail and slab dimensions for the long-term deflection
study. In total, four LWC slabs with the same mix design were internally reinforced with 4 N12
deformed steel bars. The slabs were similarly simply supported having span length of 3500 mm,
depth 161 mm and width 400 mm. The side and bottom cover of concrete are 34 and 19 mm
respectively. The steel bars are set in equal spacing through the slab width.

Figure 9.1: Cross section and reinforcing detail of LWC slab specimens in long-term deflection test

Figure (9.2) shows a general view of the test setup to monitor the time-dependent behaviour
of the four LWC slabs. The figure also shows the support condition at two edges of each slab,
the pads under the concrete loading blocks, the cables connected to the embedded strain gauges

346
Chapter 9 – Flexural long-term deflection

and the data taker to record the strain variation of the steel bars and concrete surface. The strain
data and flexural deflection were recorded in 60 minutes intervals during the test period.

Figure 9.2: General view of the long-term deflection test set up of LWC slabs

The slabs were divided into two groups to apply the service loads. The gravity load on each
pair of slab specimens was different. The sustained loads were consisted of the self-weight plus
superimposed sustained loads via carefully constructed and arranged concrete blocks supported
on the top surface of the specimens.

9.2.2. Monitoring points of time-dependent flexural deflection

The time-dependent deflection of slabs was recorded in two different points along the span
length. As shown in Figure (9.2), the deflection at mid-span as the maximum flexural deflection
was recorded by Linear Variable Differential Transformer (LVDT). Another LVDT at the third
point of the slab recorded the deflection of the slab at the boundary of the middle third high
moment region. Comparing the recorded deflection at these two points provide an overview of
the changes in the curvature of slab under the uniformly distributed flexural loading. It can also
give an insight to the time-dependent variation of section stiffness along the span.

347
Chapter 9 – Flexural long-term deflection

9.2.3. Time-dependent Deflection of Lightweight Concrete Slabs

The time-dependent deflection was monitored under the service load and the effects of creep
and shrinkage. For laboratory investigation of long-term deflection of slabs, a total of four
singly reinforced LWC one-way slab specimens with identical length, cross-section and
reinforcing details shown in Figure (9.2), were monitored for up to 550 days to measure the
time-dependent development of strain and deformations under service loads.

The slabs were cast and moist for 14 days and then subjected to uniformly distributed loads at
the age of 14 days. The steel strains within the high-moment regions in all slabs, the concrete
surface strains at the tensile steel level, the concrete compressive strain at the top surface,
deflection at the mid-span and third point of span length and crack spacing were recorded all
over the testing period. The required properties of LWC including the compressive and tensile
strength were measured on companion specimens in forms of the cylinder, cube and prism and
unreinforced blocks at various ages. Also the elastic modulus, creep coefficient and free
shrinkage in the concrete specimens were recorded. The measured mechanical properties are
presented and discussed in Chapter 5 of this study.

The applied service load in the LWC slab specimens were different gravity loads, consisting
of self-weight plus sustained superimposed loads via carefully constructed and arranged
concrete blocks supported from the top (of the specimens). To study the influence of loading
history on the time-dependent deflection of slabs, two different levels of service loads at the age
of 14 days were applied to the slabs. To provide the sustained loading, rectangular concrete
blocks of predetermined size and weight were cast and weighed prior to the commencement of
the test. As shown in Figure (9.3), the loading blocks were arranged in equal distances on the
top surface of each specimen to achieve the desired uniformly distributed sustained load level.

Figure 9.3: Loading pads and concrete block on the slabs for time-dependent testing
348
Chapter 9 – Flexural long-term deflection

Two sustained load levels were considered, namely 40% and 30% of the ultimate design load,
and designated as load conditions ‘a’ and ‘b’, respectively. The slab specimens were subjected
to uniformly distributed sustained loads, UDL + self-weight. All measurements were taken
within the high moment region, i.e. the middle third of the span where M ≥ 90% M max. For long-
term tests the loading arrangement and high moment regions are shown in Figure (9.4).

9.2.4. Testing Duration

The time-dependent deflection of LWC slabs was monitored up to 550 days after loading at
the age of 14 days. In this period, the changes of strain in the steel bars and side and top surface
of the concrete were recorded also. The deflection and strain data from the first instance of
loading were recorded every minute for up to 24 hours. Then the data taker recorded data every
hour until the end of the test program.

Figure 9.4: Illustrative sustained loads slab specimens

9.2.5. Experimental Data

The flexural deflection of LWC slabs was monitored and recorded immediately after loading
and through the test duration by two LVDTs installed at the monitoring points. Measuring the
flexural deflection at mid-span and the boundary point of the high moment region moment in
Figure (9.4) provides a comparable data of deflection changes along the span length.
Considering different levels of cracking along the span length due to the bending moment and
effects of creep and shrinkage over a long period of time, the recorded deflection can give an
insight to the varying moment of inertia along the span.

Table (9.1) presents the detailed information of the loading level, service and ultimate moments
of the LWC slabs.

349
Chapter 9 – Flexural long-term deflection

Table 9.1: loading level and bending capacity of the LWC slabs in the long-term test

Wa Ma Mu (KN.m) Ma / Mu (%)
Slab
(KN/m) (KN.m) 14 day 28 day 14 day 28 day
LWC-A 7 10.72 27.4 27.65 39.13 38.78

LWC-B 7 10.72 27.4 27.65 39.13 38.78

LWC-C 5.29 8.095 27.4 27.65 29.55 29.28

LWC-D 5.29 8.095 27.4 27.65 29.55 29.28

9.2.6. Time-Dependent Deflection at Monitoring Points

Figures (9.5) to (9.8) show the measured time-dependent deflection of LWC slabs under the
sustained service loads and effects of creep and shrinkage for 550 days. The Slabs LWC-A and
LWC-B are subjected to loading equal to 40% of their ultimate bending capacity, while slabs
LWC-C and LWC-D are loaded by 30% of the ultimate bending capacity of the section at the
critical point.

The loadings equal to 30% and 40% of the ultimate bending moment are considered as LL1
and LL2 in this chapter respectively. The index “1” and “2” after the slab identity indicates the
deflection at mid-span and the point on the border the high moment region respectively. For
example, LWC-A1 denotes the deflection at mid-span of slab LWC-A. The diagrams in Figures
(9.5) to (9.8) consist of the initial deflection of slabs at the time of loading (instantaneous
deflection)

40
35
30
Deflection (mm)

25
20
15
10 LWC-A1

5 LWC-A2

0
0 50 100 150 200 250 300 350 400 450 500 550
Age (days)

Figure 9.5: Time-dependent deflection of slab LWC-A under LL2 at monitoring points

350
Chapter 9 – Flexural long-term deflection

40
35
30
Deflection (mm)
25
20 LWC-B1
15 LWC-B2
10
5
0
0 50 100 150 200 250 300 350 400 450 500 550
Age (days)
Figure 9.6: Time-dependent deflection of slab LWC-B under LL2 at monitoring points

40
35
30
Deflection (mm)

25
20
15 LWC-C1
10 LWC-C2

5
0
0 50 100 150 200 250 300 350 400 450 500 550
Age (Days)
Figure 9.7: time-dependent deflection variation of slab LWC-C under LL1 at monitoring points

351
Chapter 9 – Flexural long-term deflection

40
35
30
Deflection (mm)
25
20
LWC-D1
15
LWC-D2
10
5
0
0 50 100 150 200 250 300 350 400 450 500 550
Age (days)
Figure 9.8: Time-dependent deflection variation of slab LWC-D at monitoring points

Figure (9.9) compares the corresponding time-dependent deflection of slabs at mid-span


subjected to two distinct loading levels of LL1 and LL2. Figure (9.10) compares the
corresponding time-dependent deflection of slabs at the third point of slabs under two different
loading levels. The diagrams in Figure (9.9) consist of the initial deflection of slab at the time
of loading (instantaneous deflection)

40
35
30
Deflection (mm)

25
20
15 LWC-A1

10 LWC-B1
LWC-C1
5
LWC-D1
0
0 50 100 150 200 250 300 350 400 450 500 550
Time (Days)

Figure 9.9: Time-dependent deflection of LWC-A and LWC-B slabs under LL2 and LWC-C and LWC-D
under LL1 at mid-span

352
Chapter 9 – Flexural long-term deflection

40
35
30
Deflection (mm)
25
20
LWC-A2
15
LWC-B2
10 LWC-C2
5 LWC-D2
0
0 50 100 150 200 250 300 350 400 450 500 550
Age (days)
Figure 9.10: Time-dependent deflection of LWC-A and LWC-B slabs under LL2 and LWC-C and LWC-D
under LL1 at the border of high moment region

The general trend of the time-dependent deflection of slabs at two monitoring points during
the monitoring time is similar. However, the ratio of deflection at mid-span to the deflection at
the border point of high moment region is considerably difference in the slabs.

Figure (9.11) compares the above-mentioned deflection ratio in the monitoring points of
LWC slabs. It can be concluded that the higher the loading level, the higher the deflection ratio
in all slabs. In other words, the higher loading level causes more primary cracking in the critical
section of the slab that in turn, increases the time-dependent deflection. While, due to change in
the primary cracking pattern of the section at the point on the border of high moment region,
variation of the deflection ratio with time is not considerable.

353
Chapter 9 – Flexural long-term deflection

1.2

Deflection ratio 1.15

1.1
LWC-A1 / LWC-A2
LWC-B1 / LWC-B2
1.05 LWC-C1 / LWC-C2
LWC-D1 / LWC-D2
1
0 50 100 150 200 250 300 350 400 450 500 550
Age (days)

Figure 9.11: Ratio of deflection at mid-span to deflection at border point of high moment region under LL1
(slabs LWC-C and LWC-D) and LL2 (slabs LWC-A and LWC0B)

Figure (9.12) compares the ratio of deflection of corresponding points in the slabs under two
distinct loading levels. Slabs LWC-A and LWC-C are considered as the representative of slabs
under higher and lower loading levels respectively. Despite differences at the initial stage of
flexural deformation, the general trend of deflection variation with time is similar in the
corresponding points of slabs under different loading levels.

1.8
LWC-A1 / LWC-C1
Deflection ratio

LWC-A2 / LWC-C2
1.6

1.4

1.2
0 50 100 150 200 250 300 350 400 450 500 550
Age (days)

Figure 9.12: Ratios of deflection at corresponding points of LWC-A under LL2 and LWC-C under LL1

354
Chapter 9 – Flexural long-term deflection

9.3. RATIO OF LONG-TERM TO INSTANTANEOUS DEFLECTION

The long-term deflection is considered as multiplier of the instantaneous deflection in


majority of the design codes. Table (9.2) presents the ratio of long-term deflection to the
instantaneous deflection of LWC slabs at different ages.

Table 9.2: Ratio of total to instantaneous deflection of LWC slabs at different ages

Age (days)

Slab 14 21 28 56 63 100 128 156 200 249 300 349 400 449 500 532

LWC-A 1 1.37 1.54 1.86 1.90 1.97 2.11 2.17 2.32 2.48 2.68 2.84 2.89 2.87 2.90 2.89

LWC-B 1 1.31 1.47 1.78 1.82 1.88 2.01 2.07 2.22 2.37 2.57 2.73 2.78 2.77 2.79 2.79

LWC-C 1 1.53 1.81 2.35 2.43 2.53 2.78 2.88 3.14 3.43 3.79 4.09 4.18 4.16 4.20 4.19

LWC-D 1 1.65 1.94 2.29 2.34 2.50 2.74 2.80 3.08 3.36 3.70 4.02 4.09 4.08 4.12 4.10

Regardless of the loading level, the deflection multiplier in design codes varies in the range
of 2 to 3. According to Table (9.2), this ratio is in agreement with the long-term deflection of
slabs at higher loading level. By 10% increase in the loading level, the deflection multiplier
increases about 50% above the range recommended in most design codes.

Figures (9.13) and (9.14) graphically compare the increment of the ratio of long-term to
instantaneous deflection of LWC slabs. In Figure (9.13), LWC-A1 and LWC-B1 are
representatives for the ratio of long-term to instantaneous deflection of slabs LWC-A and LWC-
B at mid-span of slab subjected to LL2 (0.4Mu), respectively.

In Figure (9.14), LWC-C1 and LWC-D1 are representatives for the ratio of long-term to
instantaneous deflection of slabs LWC-C and LWC-D at mid-span of slab subjected to LL1
(0.3Mu), respectively.

355
Chapter 9 – Flexural long-term deflection

4.5

3.5

Δ-long-term / Δ-inst ratio


3

2.5

2
LWC-A1
1.5

1 LWC-B1

0.5

0
0 50 100 150 200 250 300 350 400 450 500 550
Age (days)
Figure 9.13: Ratio of long-term to instantaneous deflection in LWC slabs under LL2 (0.4 Mu )

4.5

3.5
Δ-long-term / Δ-inst ratio

2.5

2 LWC-C1
1.5
LWC-D1
1

0.5

0
0 50 100 150 200 250 300 350 400 450 500 550
Age (days)
Figure 9.14: Ratio of long-term to instantaneous deflection in LWC slabs underLL1 (0.3Mu)

The ratio of long-term to instantaneous deflection of LWC slabs changes by time. The ratio
changes from 1 in the first loading time to 2.89 after 525 days sustained loading in the slabs
subjected to LL2. For the slabs subjected to LL1, this ratio changes from 1 at the loading day to
4.19 after 525 sustained d loading.

The instantaneous and final deflection after 525 days in both monitoring points of the slabs
LWC-A and LWC-B are greater the corresponding deflections in LWC-C and LWC-D. Despite
the greater values of the instantaneous deflection in the slabs under LL2, the relative increment
of the long-term deflection in the slabs subjected to lower sustained loading is considerably

356
Chapter 9 – Flexural long-term deflection

high. Considering the constant value of the sustained load of slabs in the monitoring period, it
shows the significant effect of the creep and shrinkage on the long-term deflection of LWC
slabs.

9.4. RATE OF TIME DEPENDENT DEFLECTION IN LIGHTWEIGHT


CONCRETE SLABS

Not only the total deflection after a long time is a crucial parameter in design of the
reinforced concrete structures, but also the rate of deflection increment with time is important. A
relatively rapid deflection rate limits the service life and consequently applies more conservative
design considerations. However, in a concrete structure with slow increment of the time-
dependent deflection, service life is not significantly affected by the deflection rate.

Figures (9.15) and (9.16) show the rate of time-dependent deflection of LWC slabs. In Figure
(8.15), the rate of deflection increment in the slabs LWC-C and LWC-D at the mid-span is
presented. Both slabs are subjected to the lower loading level, LL1. Time-dependent deflection
of these slabs is sharply growing after the first loading until about 50 days. The deflection rate is
changing from 8mm per day in the first days to less than 1 mm per day after 50 days. This rate
of time-dependent deflections tends to a constant rate at about after 150 days until 525 days.
Despite the tendency of deflection rate to zero after 525, the time-dependent deflection never
stops in the lightweight concrete slabs.

In Figure (9.16), the rate of deflection increment in the slabs LWC-A and LWC-B at the mid-
span is presented. Both slabs are subjected to the higher loading level, LL2. Time-dependent
deflection of these slabs is abruptly increasing after the first loading until about 50 days. The
deflection rate is changing from 8mm per day in the first days to less than 1 mm per day after 50
days. This rate of time-dependent deflections tends to a constant rate at about after 150 days
until 525 days. Despite the tendency of deflection rate to zero after 525, the time-dependent
deflection continues even by very low rate.

357
Chapter 9 – Flexural long-term deflection

14
13
12
11

Deflection rate (mm/ day)


10
9
8 LWC-C LWC-D
7
6
5
4
3
2
1
0
0 50 100 150 200 250 300 350 400 450 500 550
Age (days)

Figure 9.15: Deflection rate of slabs C and D under LL1 (0.3Mu)

14
13
12
11
Deflection rate (mm/ day)

10
9
8 LWC-A LWC-B
7
6
5
4
3
2
1
0
0 50 100 150 200 250 300 350 400 450 500 550
Age (days)

Figure 9.16: Deflection rate of slabs A and B under LL2 (0.4Mu)

9.5. CRACKING HISTORY OF THE LIGHTWEIGHT CONCRETE


SLABS AT EARLY AGES OF LOADING

According to Figures (9.15) and (9.16), the highest rate of the time-dependent deflection in
the monitored LWC slabs happened before the age of 50 days of loading. In other words,
majority of the time-dependent deflection of slabs under two different loading levels happened
before 50 days of sustained loading.

The load increment changes the cracking pattern in terms of the crack width and crack
spacing. The more cracks, the higher the effects of creep and shrinkage on the time-dependent
deflection of slabs and vice versa. Figures (9.17) to (9.20) show the cracking distribution of
358
Chapter 9 – Flexural long-term deflection

LWC slabs at two ages, one week after loading and 6 weeks (about 50 days) after loading,
respectively.

Figure 9.17: Cracking pattern of slab LWC-A at 7 and 48 days after sustained loading

Figure 9.18: Cracking pattern of slab LWC-B at 7 and 48 days after sustained loading

Figure 9.19: Cracking history of slab C at 7 and 48 days after sustained loading

Figure 9.20: Cracking history of slab D at 7 and 48 days after sustained loading

359
Chapter 9 – Flexural long-term deflection

9.6. SIMIALR STUDIES ON THE CONVENTIONAL AND SELF-


COMPACTING CONCRETE SLABS

This study compares the experimental results of the instantaneous and time-dependent
deflection of three different types of concrete slabs. The experimental program consists of six
one-way slabs including three pairs of SCC slabs conducted by Aslani (2014) [21], LWC slabs
in this research program, and CC slabs conducted by Nejadi (2005) [22] with different mixture
designs.

The slabs were simply supported with identical span length, section dimensions and
reinforcement arrangement. The standard size of the cylindrical specimen (150 ¯300 mm) was
used to measure the compressive strength, modulus of elasticity and creep coefficient of the
mixtures. Also, the shrinkage strain was measured by the standard prisms in laboratory
controlled conditions. The test specimens were kept covered in a controlled chamber at 23 ± 2oC
for 24 hours until demoulding. The relative humidity of the environment was maintained at
50±5% during the tests. Then, the specimens were placed in water pre-saturated with lime at
20oC. The compressive strength test was performed according to AS-1012.14 (1991) [23] and
ASTM-C39 (2000) [24] instructions, and the loading rate of the machine on the cylinders was
0.3 MPa/s until failure. The modulus of elasticity test was performed according to the
instructions in AS-1012.17 (2014) [25] and ASTM-C469 (2000) [26].

9.6.1. Conventional Concrete Slabs

Figure (9.21) shows the experiment conducted by Nejadi (2005) on the conventional concrete
slabs. The time-dependent deflection of slabs at mid-span was monitored for 400 days in the
laboratory conditions.

One batch of commercially pre-mixed concrete was used to manufacture all the test and
companion specimens. Two sustained load levels were considered, being 30% and 20% of the
ultimate design load, and designated load conditions ‘a’ and ‘b’, respectively. The applied load
consisted of self-weight plus superimposed sustained loads at the age of 14 days via carefully
constructed and arranged concrete blocks supported at the top (of the specimens).

360
Chapter 9 – Flexural long-term deflection

Figure 9.21: The loading of convention concrete slabs for time-dependent deflection monitoring

The summary of the long-term deflection results for the CC slabs is shown in Figure (9.22).

30

25
Deflection (mm)

20

15

10 S3-A
S3-B
5

0
0 50 100 150 200 250 300 350 400
Age (days)
Figure 9.22: Time-dependent deflection of slabs at mid-span (Nejadi, 2005)

9.6.2. Self-Compacting Concrete Slabs

In an experiment by Aslani (2014), a total of eight SCC and Fibre Reinforced Self-
Compacting Concrete (FRSCC) slabs with the same cross-section and details as for the short-
361
Chapter 9 – Flexural long-term deflection

term tests were monitored for up to 240 days to measure the time-dependent development of
cracking and deformations under service loads. Deflections at mid-span, crack widths, crack
patterns, steel strains within the high moment region, and concrete surface strains at the steel
level were recorded under the sustained service load and effects of creep and shrinkage. For
this purpose, four SCC mixes – two plain SCC, two steel, two polypropylene, and two hybrid
FRSCC slab specimens – are considered in the test program.

Two sustained load levels were considered, being 40% and 30% of the ultimate design load,
and designated load conditions ‘a’ and ‘b’, respectively. Figure (9.23) shows the general view
of the test arrangement for time-dependent deflection monitoring of self-compacting concrete
slabs.

Figure 9.23: The loading of Self-compacting concrete slabs for time-dependent deflection monitoring (Aslani,
2014)

Summary of the long-term deflection results for SCC and FRSCC slabs under higher and
lower loading levels are shown in Figures (9.24) and (9.25) respectively.

35

30

25
Deflection (mm)

20

15

10
D-SCC-A DS-SCC-A
5
N-SCC-A S-SCC-A

0
0 50 100 150 200 250
Age (days)

Figure 9.24: Time-dependent deflection of SCC and FRSCC slabs under 0.4Mu loading (Aslani, 2014)

362
Chapter 9 – Flexural long-term deflection

35
D-SCC-B DS-SCC-B
30
N-SCC-B S-SCC-B
25

Deflection (mm)
20

15

10

0
0 50 100 150 200 250
Age (days)

Figure 9.25: Time-dependent deflection of SCC and FRSCC slabs under 0.3Mu loading (Aslani, 2014)

9.7. COMPARISON OF TIME-DEPENDENT DEFLECTION IN


LIGHTWEIGHT, SELF-COMPACTING AND CONVENTIOANL
CONCRETE SLABS

Table (9.3) summarizes the ultimate moment capacity (Mu), applied service moment at the
critical section (Ma), predicted elastic deflection (Δe), and the recorded instantaneous deflection
(Δins) immediately after the load application on the top surface of simply supported slabs.

The monitoring period of long-term deflection for LWC, SCC and LWC slabs are 550, 250
and 400 days respectively. Therefore, the long-term deflection of the LWC, SCC and CC slabs
are compared for 250 days. However, the LWC and CC slabs are compared for 400 days.

Figure (9.26) compares the mid-span deflection versus time in the CC, LWC and SCC slabs
under the higher loading level. As expected from the measured mechanical properties (see
chapter 5), the CC slab shows the leading ratio of the total to instantaneous deflection, while the
SCC slab is placed in the lower limit of the diagram. The variation of the ratio in LWC and CC
slab is similar up to 60 days. Then, the deflection of the CC slab increases with a higher rate
until the end of the monitoring period. The time-deflection diagram in the SCC slab follows a
relatively constant increasing rate from the first loading at the age of 14 days.

363
Chapter 9 – Flexural long-term deflection

Table 9.3: Loading level, predicted elastic deflection and recorded instantaneous deflection of slabs

Δ ins
Wa Mu (KN.m) Ma Ma / Mu (%) Δ e (mm)
Slab
(KN/m) 14 day 28 day (KN.m) 14 day 28 day 14 day 28 day (mm)

N-SCC-a 7.31 27.47 27.78 11.189 40.73 40.27 3.18 2.9 12.1

N-SCC-b 5.41 27.47 27.78 8.28 30.14 29.8 2.36 2.15 5.89

D-SCC-a 7.26 27.84 28.05 11.12 39.94 39.65 3.5 2.85 7.65

D-SCC-b 5.36 27.84 28.05 8.21 29.48 29.27 2.58 2.11 7.59

S-SCC-a 7.3 27.73 28.05 11.18 40.32 39.85 3.45 2.87 6.41

S-SCC-b 5.4 27.73 28.05 8.27 29.83 29.48 2.55 2.12 2.91

DS-SCC-a 7.33 28.05 28.34 11.23 40.03 39.63 3.29 2.85 8.98

DS-SCC-b 5.43 28.05 28.34 8.32 29.66 29.36 2.44 2.11 5.14

LWC-1 7 27.4 27.65 10.72 39.13 38.78 4.99 4.57 11.96

LWC-2 7 27.4 27.65 10.72 39.13 38.78 4.99 4.57 11.4

LWC-3 5.29 27.4 27.65 8.095 29.55 29.28 3.77 3.45 6.3

LWC-4 5.29 27.4 27.65 8.095 29.55 29.28 3.77 3.45 5.98

CC-a 4.44 26.03 27.05 6.8 26.12 25.14 2.73 2.5 11.8

CC-b 3.46 26.03 27.05 5.3 20.36 19.59 2.13 1.95 5.04

35

30

25
Deflection (mm)

20

15

10
LWC-A D-SCC-A DS-SCC-A
5
N-SCC-A S-SCC-A CC-a

0
0 50 100 150 200 250
Age (days)

Figure 9.26: Comparison of deflection vs. time in SCC, LWC and CC slabs under higher loading level

364
Chapter 9 – Flexural long-term deflection

Figure (9.27) also compares the mid-span deflection versus time in the CC, LWC and SCC
slabs under the lower loading level.

35

30 LWC-D D-SCC-B DS-SCC-B

N-SCC-B S-SCC-B CC-b


25
Deflection (mm)

20

15

10

0
0 50 100 150 200 250
Age (days)

Figure 9.27: Comparison of deflection vs. time in SCC, LWC and CC slabs under lower loading level

To evaluate the total and time-dependent deflection over a longer period, the variation of
mid-span deflection versus time in LWC and CC slabs under higher loading level is given in
Figure (9.28) for the sustained loading period of 400 days.

35

30

25
Deflection (mm)

20

15
LWC-A
10
LWC-B
5 CC-S3-A

0
0 50 100 150 200 250 300 350 400
Age (days)

Figure 9.28: Deflection vs. time in LWC and CC slabs under higher loading level

365
Chapter 9 – Flexural long-term deflection

The time-dependent deflection increases at a decreasing rate, with much of the long-term
deflection taking place in the first 100 days after first loading in all slabs of different concrete
types. However, the rate of deflection increase before and after that time is not entirely similar.

9.8. TIME-DEPENDENT DEFLECTION IN CODES OF PRACTICE

Design codes recommend simplified methods of calculating the time-dependent and long-
term deflection of concrete structures. Despite the uncertainties, it is necessary for the calculated
deflection to match the real deflection precisely. In some universally applied codes of practice,
the long-term deflection is predicted as a multiple of the instantaneous or short-term deflection.
To calculate the long-term deflection, ACI318-08 [27] and CSA.A23.04[20] recommend a
simplified procedure by multiplying the instantaneous deflection by a multiplier. AS3600-
09[19] uses a similar procedure with some modifications to calculate the time-dependent
deflection. Table (8.4) presents the equations in the above mentioned codes to calculate the
long-term deflection from instantaneous deflection.

Table 9.4: Calculation of long-term deflection in codes of practice

Code Total deflection Coefficients


ߜ
ACI-318-08[27] ο௜௡௦ ሺͳ ൅ ሻ δ: 1.2, 1.25, 1.4 for 180, 225, 365 days
ͳ ൅ ͷͲߩᇱ
ߦ
CSA.A23-04[20] ο௜௡௦ ሺͳ ൅ ሻ ζ: 1.2, 1.25, 1.4 for 180, 225, 365 days
ͳ ൅ ͷͲߩᇱ
kcs: creep and shrinkage coefficient
AS-3600-09[19] ο௜௡௦ ൅ ݇௖௦ ο௦௨௦
Δsus: time-dependent deflection under sustained load

Figures (9.29 a, b) show the predictions of AS-3600 (2009)[19] , ACI-318 (2008)[27] and
CSA.A23 (2004)[20] for the instantaneous and total (long-term) deflection of SCC, LWC and
CC slabs under two different levels of sustained loading respectively.

366
Chapter 9 – Flexural long-term deflection

35
SCC-b LWC-b CC-b
30

Predicted deflection (mm)


25

20

15
(a)
10

0
Δins Δtot Δins Δtot Δins Δtot
ACI-318 AS-3600 CSA.A23
Codes of practice
50
45 SCC-a LWC-a CC-a
Predicted deflection (mm)

40
35
30
25
20
(b)
15
10
5
0
Δins Δtot Δins Δtot Δins Δtot
ACI-318 AS-3600 CSA.A23
Codes of practice

Figure 9.29: Predicted Δins and Δtot of SCC, LWC and CC slabs under a) LL1, b) LL2

According to Figures (8.29 a, b), there is a considerable difference between the predicted
Δins-14 and Δlong of slabs under different load values. The instantaneous and total deflection in the
slabs under higher loading level is significantly greater than those of the slabs under lower
loading levels, especially by ACI-318[27] and CSA.A23[20] predictions. However, in predicted
values of AS-3600[19], the difference between the Δins-14 and Δlong is less, compared to the other
design codes.

All listed codes in Figures (9.29 a, b) predict the highest values of Δins-14 and Δlong for LWC
slabs under both loading levels, however, AS-3600[19] predicts the lowest Δins-14 and Δlong values
for SCC slabs under both loading levels.

367
Chapter 9 – Flexural long-term deflection

9.9. RATIO OF LONG-TERM TO INSTANTANEOUS DEFLECTION IN


LIGHTWEIGHT, SELF-COMPACTING AND CONVENTIONAL
CONCRETE SLABS

Since the total monitoring period for three different types of concrete slabs are not the same,
the individual recording of each concrete type is presented. Then the comparative records for
the compatible duration are illustrated. The ratio of long-term deflection of CC slabs to the
short-term deflection under sustained loading in different ages is presented in Table (9.5)
respectively.

Table 9.5: Long-term to short-term deflection ratio in slab series tested by Nejadi and Gilbert (2005)

Age (day) 14 21 28 56 60 95 122 200 285 330 394


Slab S3-a 1 1.54 1.67 1.96 2.1 2.25 2.4 2.61 2.72 2.77 2.79

Slab S3-b 1 1.96 2.19 2.69 2.96 3.21 3.5 3.98 4.35 4.45 4.54

The ratio of long-term deflection of SCC and FRSCC slabs to the short-term deflection under
sustained loading in different ages is presented in Table (9.6) respectively.

Table 9.6: Long-term to short-term deflection ratio in SCC and FRSCC slabs

Age (days)
Slab
14 21 28 45 60 95 122 200 240

N-SCC-a 1 1.18 1.28 1.42 1.55 1.73 1.72 2 2.04

N-SCC-b 1 1.42 1.63 1.91 2.14 2.48 2.65 2.98 3.07

D-SCC-a 1 1.31 1.43 1.71 1.82 2.03 2.13 2.31 2.32

D-SCC-b 1 1.28 1.4 1.67 1.75 1.94 2.03 2.2 2.21

S-SCC-a 1 1.44 1.79 2.31 2.53 2.91 3.12 3.44 3.47

S-SCC-b 1 2.15 2.66 3.94 4.53 5.52 6.06 6.91 9.65

DS-SCC-a 1 1.3 1.48 1.72 1.89 2.13 2.26 2.36 2.37

DS-SCC-b 1 1.42 1.7 2.06 2.31 2.67 2.88 3.06 3.08

The comparative presentation of the deflection ratios in Tables (9.2), (9.5) and (9.6) in 225
days duration is presented in Table (9.7).

368
Chapter 9 – Flexural long-term deflection

Table 9.7: Ratio of long-term to instantaneous deflection in LWC, CC and SCC slabs

Age (days)
Slab
14 28 49 72 100 121 149 172 200 214 225
N-SCC-A 1 1.27 1.48 1.62 1.73 1.82 1.92 1.97 2.01 2.02 2.03
N-SCC-B 1 1.61 1.96 2.26 2.47 2.64 2.8 2.9 2.98 3.01 3.02
D-SCC-A 1 1.43 1.71 1.88 2.02 2.12 2.13 2.18 2.2 2.21 2.22
D-SCC-B 1 1.38 1.65 1.82 1.94 2.01 2.12 2.17 2.2 2.22 2.23
S-SCC-A 1 1.76 2.34 2.69 2.94 3.12 3.31 3.34 3.43 3.44 3.45
S-SCC-B 1 2.59 4.08 4.88 5.64 6.11 6.59 6.67 6.86 6.87 6.89
DS-SCC-A 1 1.47 1.78 1.97 2.15 2.24 2.3 2.35 2.36 2.37 2.38
DS-SCC-B 1 1.68 2.15 2.43 2.72 1.82 2.98 3.04 3.07 3.08 3.09

Figure (9.30) compares the mid-span deflection versus time in the CC, LWC and SCC slabs
under higher loading level.

4.5

3.5
Δ long-term / Δ inst.

2.5

1.5
CC-a
1
SCC-a
0.5
LWC-a
0
0 50 100 150 200 250
Age (days)
Figure 9.30: The ratio of long-term to instantaneous deflection under higher loading level

Figure (9.31) also compares the mid-span deflection versus time in the CC, LWC and SCC
slabs under lower loading level. Similar to the Figure (9.30), the CC slab poses the largest ratio
of the long-term to instantaneous deflection, while the SCC slab experiences the lowest ratio
over all the testing period.

369
Chapter 9 – Flexural long-term deflection

2.5

Δ long-term / Δ inst.
2

1.5

1
CC-b

0.5 SCC-b
LWC-b
0
0 50 100 150 200 250
Age (days)
Figure 9.31: The ratio of long-term to instantaneous deflection under lower loading level

The ratio of the long-term to instantaneous deflection is the result of the combined effects of
the creep and shrinkage on the member. However, only the time-dependent deflection due to
creep may be correctly expressed as a multiple of the instantaneous deflection. The shrinkage
induced deflection is dependent on the free shrinkage that is independent of the instantaneous
deflection [28]. Table (9.8) shows the free shrinkage strain of the CC, LWC and SCC specimens
at 240 days after casting. The creep coefficients of the companion specimens of the concrete
slabs since the loading age are presented in Table (9.8) also. The creep coefficients are measures
in the cylinder specimens under the corresponding sustained compression stress at slab
specimens.

Table 9.8: Free shrinkage strain and creep coefficient of CC, LWC and SCC mixtures

CC LWC SCC
Shrinkage strain (με) 784 1180 870
Creep coefficient 1.5 2.415 1.96

According to the Table (8.8), the shrinkage strain in LWC is considerably higher than that of
the CC and SCC mixtures. The free shrinkage strain in LWC and SCC mixtures is 51% and
10% greater than that of the CC mixture respectively. However, the ratio of long-term to
instantaneous deflection in LWC and SCC slabs is 90% and 75% of the corresponding
deflection ratio in CC slab under lower loading levels respectively. In the slabs subjected to

370
Chapter 9 – Flexural long-term deflection

higher loading levels, the deflection ratio in LWC and SCC slabs reaches to 79% and 73% of
the corresponding deflection ratio in CC slab respectively.

Comparing the deflection ratios in Figures (9.30), (9.31) and the free shrinkage values in
Table (9.8), the ratio of long-term to instantaneous deflection cannot be concluded from the free
shrinkage trend over the time period. Also, the rate of deflection and the free shrinkage rate in
different types of concrete are not compatible.

The creep coefficient in LWC and SCC mixtures is 1.61 and 1.3 times higher than that of the
CC mixture respectively. Although the ratio of long-term to instantaneous deflection is
frequently used as the deflection multiplier in the codes of practice, the diagrams and data
presented in Figures (9.30), (9.31) and Table (9.8) are not thorough enough to establish any
relationship. In other words, the creep coefficient is not a good indicator of the time-dependent
deflection increment in different types of concrete, without another mechanism of assessment.
The loading value, type of concrete and mixture components, loading age, environmental effects
and testing equipment strongly affect the time-dependent and long-term deflection of slabs.

The long-term deflection after 240 days in LWC slabs is 0.99 and 0.92 of the corresponding
deflection in CC slabs under the lower and higher loading levels respectively. Also, the
deflection of SCC slabs under the lower and upper loading levels is 0.85 and 0.78 of the
deflection under corresponding loading levels in CC slabs.

Figure (9.32) compares the predicted and experimental ratio of Δlong / Δins-14 of slabs under two
loading levels. The experimental ratio ranges from 2.04 to 4.35, however, it changes in the
range of 1.89 to 2.31 in the predictions of codes of practice.

According to Figure (8.32), AS-3600[19] gives the lowest ratio of Δlong / Δins-14 of all slabs.
Except for the SCC-b slab, the experimental ratio in all slabs is higher than the predicted values
of design codes. The difference between the experimental and predicted ratio of the deflection
in slabs increases by increasing the loading value.

The CC slabs, particularly CC-a, show the highest difference between the experimental and
predicted deflection ratio. In other words, the codes of practice give the farthest prediction of
the total to instantaneous deflection ratio in CC slabs under both loading levels.

371
Chapter 9 – Flexural long-term deflection

5
4.5 ACI-318 AS-3600
4 CSA.A23 Experimental
3.5
Δtot / Δins ratio 3
2.5
2
1.5
1
0.5
0
SCC-a SCC-b LWC-a LWC-b CC-a CC-b
Slabs
Figure 9.32: Comparing the experimental and predicted ratio of the total to instantaneous deflection

Figures (9.33) and (9.34) compare the deflection ratio in SCC and FRSCC slabs under two
different loading levels.

3.5
Δ-long-term / Δ-inst ratio

2.5

1.5
N-SCC-A D-SCC-A
1
DS-SCC-A S-SCC-A
0.5

0
0 50 100 150 200 250
Age (days)

Figure 9.33: The ratio of long-term to instantaneous deflection of SCC and FRSCC slabs (higher loading level)

372
Chapter 9 – Flexural long-term deflection

8
N-SCC-B D-SCC-B
7 DS-SCC-B S-SCC-B

Δ-long-term / Δ-inst ratio


6

0
0 50 100 150 200 250
Age (days)

Figure 9.34: The ratio of long-term to instantaneous deflection of SCC and FRSCC slabs (lower loading level)

Figure (9.35) compares the LWC, SCC and CC slabs in their real monitored period under the
higher loading level.

3.5

3
Δ-long-term / Δ-inst ratio

2.5

1.5

1
LWC-A S-SCC-A N-SCC-A
0.5
D-SCC-A DS-SCC-A CC-S3-A
0
0 50 100 150 200 250 300 350 400 450 500 550
Age (days)

Figure 9.35: The ratio of long-term to instantaneous deflection of all slabs (higher loading level)

373
Chapter 9 – Flexural long-term deflection

9.10. DEFLECTION RATE IN LIGHTWEIGHT, SELF-COMPACTING


AND CONVENTIONAL CONCRETE SLABS

Figures (9.36) and (9.37) compare the rate of deflection in the LWC, SCC, FRSCC and CC
slabs under two different loading levels respectively.

16

14

12
Deflection rate (mm/ day)

LWC-A LWC-B D-SCC-A


10
DS-SCC-A N-SCC-A S-SCC-A
8 CC-S3-A

0
0 50 100 150 200 250
Age (days)

Figure 9.36: Deflection rate of all slabs under higher loading level

7
Deflection rate (mm/ day)

6
LWC-C LWC-D D-SCC-B
5
DS-SCC-B N-SCC-B S-SCC-B
4 CC-S3-B

0
0 50 100 150 200 250
Age (days)

Figure 9.37: Deflection rate of all slabs under lower loading level

374
Chapter 9 – Flexural long-term deflection

9.11. SUMMARY

Time-dependent deflection of lightweight concrete slabs under two different levels of


sustained service loads in the experimental program is described in details. The obtained results
are analysed and discussed broadly and compared with the previously performed experimental
results by Nejadi (2005) and Aslani (2014).

Figure (9.38) shows the section details, span length and support conditions of the lightweight
concrete slabs subjected to the sustained loading for 550 days. The slabs were loaded at age 14
days after casting. Deflection of the slabs was measured at two points; the mid-span and the
point on border of the high moment region.

The loadings equal to 30% and 40% of the ultimate bending moment are considered as LL1
and LL2 in this chapter respectively. Table (9.9) presents the detailed information of the loading
level, service and ultimate moments of the LWC slabs.

Figure 9.38: Cross section and reinforcing detail of LWC slab specimens in long-term deflection test

Table 9.9: loading level and bending capacity of the LWC slabs in the long-term test

Wa Ma Mu (KN.m) Ma / Mu (%)
Slab
(KN/m) (KN.m) 14 day 28 day 14 day 28 day
LWC-A 7 10.72 27.4 27.65 39.13 38.78

LWC-B 7 10.72 27.4 27.65 39.13 38.78

LWC-C 5.29 8.095 27.4 27.65 29.55 29.28

LWC-D 5.29 8.095 27.4 27.65 29.55 29.28

375
Chapter 9 – Flexural long-term deflection

Figure (9.39) compares the corresponding time-dependent deflection of slabs at mid-span


subjected to two distinct loading levels of LL1 and LL2. Figure (9.40) compares the
corresponding time-dependent deflection of slabs at the point on the border of high moment
region of slabs under two different loading levels. The diagrams in Figures (9.39) and (9.40)
consist of the initial deflection of slab at the time of loading (instantaneous deflection)

40
35
30
Deflection (mm)

25
20
15 LWC-A1

10 LWC-B1
LWC-C1
5
LWC-D1
0
0 50 100 150 200 250 300 350 400 450 500 550
Time (Days)
Figure 9.39: Time-dependent deflection of LWC-A and LWC-B slabs under LL2 and LWC-C and LWC-D
under LL1 at mid-span

40
35
30
Deflection (mm)

25
20
LWC-A2
15
LWC-B2
10 LWC-C2
5 LWC-D2
0
0 50 100 150 200 250 300 350 400 450 500 550
Age (days)
Figure 9.40: Time-dependent deflection of LWC-A and LWC-B slabs under LL2 and LWC-C and LWC-D
under LL1 at the border of high moment region

The general trend of the time-dependent deflection of slabs at two monitoring points during
the monitoring time is similar. However, the ratio of deflection at mid-span to the deflection at
the border point of high moment region is considerably difference in the slabs.

376
Chapter 9 – Flexural long-term deflection

The long-term deflection is considered as multiplier of the instantaneous deflection in


majority of the design codes. Table (9.10) shows the ratio of long-term to instantaneous
deflection in the LWC slabs under two different loading levels.

Table 9.10: Ratio of total to instantaneous deflection of LWC slabs at different ages

Age (days)

Slab 14 21 28 56 63 100 128 156 200 249 300 349 400 449 500 532

LWC-A 1 1.37 1.54 1.86 1.90 1.97 2.11 2.17 2.32 2.48 2.68 2.84 2.89 2.87 2.90 2.89

LWC-B 1 1.31 1.47 1.78 1.82 1.88 2.01 2.07 2.22 2.37 2.57 2.73 2.78 2.77 2.79 2.79

LWC-C 1 1.53 1.81 2.35 2.43 2.53 2.78 2.88 3.14 3.43 3.79 4.09 4.18 4.16 4.20 4.19

LWC-D 1 1.65 1.94 2.29 2.34 2.50 2.74 2.80 3.08 3.36 3.70 4.02 4.09 4.08 4.12 4.10

Regardless of the loading level, the deflection multiplier in design codes varies in the range
of 2 to 3. According to Table (9.10), this ratio is in agreement with the long-term deflection of
slabs at lower loading level. By 10% increase in the loading level, the deflection multiplier
increase about 50% above the range of design codes.

- Comparing the results with previously conducted experiments on conventional and


self-compacting slabs

Table (9.11) shows the measured deflections of lightweight, self-compacting and


conventional concrete slabs at the time 0 after loading (instantaneous deflection) and the final
deflection after long-term sustained loading. As presented in Table (9.11) the test duration is not
equal in the sabs with different concrete types. The ratio of long-term to instantaneous
deflection is a crucial parameter in design of the reinforced concrete flexural elements.

Figure (9.41) compares the mid-span deflection versus time in the CC, LWC and SCC slabs
under the higher loading level. As expected from the measured mechanical properties (see
Chapter 6), the CC slab shows the leading ratio of the total to instantaneous deflection, while the
SCC slab is placed in the lower limit of the diagram. The variation of the ratio in LWC and CC
slab is similar up to 60 days. Then, the deflection of the CC slab increases with a higher rate
until the end of the monitoring period. The time-deflection diagram in the SCC slab follows a
relatively constant increasing rate from the first loading at the age of 14 days.

377
Chapter 9 – Flexural long-term deflection

Table 9.11: Ratio of the long-term to instantaneous deflection of LWC, SCC and CC slabs

Ma / Mu (%) Test duration Δ tot Δ ins Δ tot


Slab
14 day 28 day (days) (mm) (mm) / Δ ins
N-SCC-a 40.73 40.27 240 24.76 12.1 2.04
N-SCC-b 30.14 29.8 240 18.08 5.89 3.07
D-SCC-a 39.94 39.65 240 17.76 7.65 2.32
D-SCC-b 29.48 29.27 240 16.78 7.59 2.21
S-SCC-a 40.32 39.85 240 22.27 6.41 3.47
S-SCC-b 29.83 29.48 240 20.25 2.91 6.95
DS-SCC-a 40.03 39.63 240 21.31 8.98 2.37
DS-SCC-b 29.66 29.36 240 15.83 5.14 3.08
LWC-1 39.13 38.78 525 34.56 11.96 2.89
LWC-2 39.13 38.78 525 31.81 11.4 2.79
LWC-3 29.55 29.28 525 26.4 6.3 4.19
LWC-4 29.55 29.28 525 24.52 5.98 4.1
CC-a 26.12 25.14 400 32.1 11.8 2.72
CC-b 20.36 19.59 400 21.92 5.04 4.35

30

25
Deflection (mm)

20

15

10
LWC-A LWC-B D-SCC-A

5 DS-SCC-A N-SCC-A S-SCC-A


CC-S3-A

0
0 50 100 150 200 250
Age (days)
Figure 9.41: Comparison of deflection vs. time in SCC, LWC and CC slabs under higher loading level

Figure (9.42) also compares the mid-span deflection versus time in the CC, LWC and SCC
slabs under the lower loading level.

378
Chapter 9 – Flexural long-term deflection

30

25
Deflection (mm)
20

15

10 LWC-D D-SCC-B DS-SCC-B

N-SCC-B S-SCC-B CC-B


5

0
0 50 100 150 200 250
Age (days)
Figure 9.42: Comparison of deflection vs. time in SCC, LWC and CC slabs under lower loading level

Figure (9.43) compares the mid-span deflection versus time in the CC, LWC and SCC slabs
under higher loading level.

4.5

3.5
Δ long-term / Δ inst.

2.5

1.5
CC-a
1
SCC-a
0.5
LWC-a
0
0 50 100 150 200 250
Age (days)
Figure 9.43 The ratio of long-term to instantaneous deflection under higher loading level

Figure (9.44) also compares the mid-span deflection versus time in the CC, LWC and SCC
slabs under lower loading level. Similar to the Figure (9.43), the CC slab poses the largest ratio

379
Chapter 9 – Flexural long-term deflection

of the long-term to instantaneous deflection, while the SCC slab experiences the lowest ratio
over all the testing period.

Δ long-term / Δ inst. 2.5

1.5

1
CC-b

0.5 SCC-b
LWC-b
0
0 50 100 150 200 250
Age (days)
Figure 9.44: The ratio of long-term to instantaneous deflection under lower loading level

Table (9.12) shows the free shrinkage strain of the CC, LWC and SCC specimens at 240 days
after casting. The creep coefficients of the companion specimens of the concrete slabs since the
loading age are presented in Table (8.8) also.

Table 9.12: Free shrinkage strain and creep coefficient of CC, LWC and SCC mixtures

CC LWC SCC
Shrinkage strain (με) 784 1180 870
Creep coefficient 1.5 2.415 1.96

Comparing the deflection ratios in Figures (9.43), (9.44) and the free shrinkage values in
Table (9.12), the ratio of long-term to instantaneous deflection cannot be concluded from the
free shrinkage trend over the time period. Also, the rate of deflection and the free shrinkage rate
in different types of concrete are not compatible.

Figure (9.45) compares the predicted and experimental ratio of Δlong / Δins-14 of slabs under two
loading levels. The experimental ratio ranges from 2.04 to 4.35, however, it changes in the
range of 1.89 to 2.31 in the predictions of codes of practice.

380
Chapter 9 – Flexural long-term deflection

5
4.5 ACI-318 AS-3600
4 CSA.A23 Experimental
3.5
Δtot / Δins ratio 3
2.5
2
1.5
1
0.5
0
SCC-a SCC-b LWC-a LWC-b CC-a CC-b
Slabs
Figure 9.45: Comparing the experimental and predicted ratio of the total to instantaneous deflection

Comparison of the long-term to instantaneous deflection ratio with multipliers in codes of


practice

Figures (9.46 a, b) show the predictions of AS-3600 (2009)[19] , ACI-318 (2008)[27] and
CSA.A23 (2004)[20] for the instantaneous and total (long-term) deflection of SCC, LWC and
CC slabs under two different levels of sustained loading respectively.

According to Figures (9.46 a, b), there is a considerable difference between the predicted
Δins-14 and Δlong of slabs under different load values. The instantaneous and total deflection in the
slabs under higher loading level is significantly greater than those of the slabs under lower
loading levels, especially by ACI-318[27] and CSA.A23[20] predictions. However, in predicted
values of AS-3600[19], the difference between the Δins-14 and Δlong is less, compared to the other
design codes.

All listed codes in Figures (8.29 a, b) predict the highest values of Δ ins-14 and Δlong for LWC
slabs under both loading levels, however, AS-3600[19] predicts the lowest Δins-14 and Δlong
values for SCC slabs under both loading levels.

381
Chapter 9 – Flexural long-term deflection

35
SCC-b LWC-b CC-b
30
Predicted deflection (mm)

25

20

15
(a)
10

0
Δins Δtot Δins Δtot Δins Δtot
ACI-318 AS-3600 CSA.A23
Codes of practice

50
45 SCC-a LWC-a CC-a
Predicted deflection (mm)

40
35
30
25
20
(b)
15
10
5
0
Δins Δtot Δins Δtot Δins Δtot
ACI-318 AS-3600 CSA.A23
Codes of practice

Figure 9.46: Predicted Δins and Δtot of SCC, LWC and CC slabs under a)LL1, b) LL2

382
Chapter 9 – Flexural long-term deflection

REFERENCES

1. Unnikrishna Pillai, S. and D. Menon, Reinforced Concrete Design Tata Mc Graw Hill, New
Delhi, India,, 2003.
2. Vakhshouri, B. and S. Nejadi. Limitations and Uncertainties in the Long Term Deflection
Calculation of Concrete Structures. in Second International Conference on Vulnerability and
Risk Analysis and Management (ICVRAM) and the Sixth International Symposium on
Uncertainty, Modeling, and Analysis (ISUMA). 2014.
3. ACI-209.2R-08, Guide for Modeling and Calculating Shrinkage and Creep in hardened
concrete. American Concrete Institute (ACI), ACI Committee 209, Farmington Hills, MI 48331,
U.S.A, 2008.
4. ACI-318-05, Building code requirements for structural concrete ACI Committee, Standard ACI
318. American Concrete Institute, Farmington Hills, Mich, 2005.
5. Park, H., et al., Creep and effective stiffness of early age concrete slabs. 2010.
6. Park, H.-G., et al., Immediate and Long-Term Deflections of Reinforced Concrete Slabs Affected
by Early-Age Loading and Low Temperature. ACI Structural Journal, 2012. 109(3).
7. Lee, J., A. Scanlon, and M. Scanlon, Effect of Early Age Loading on Time-Dependent Deflection
and Shrinkage Restraint Cracking of Slabs with Low Reinforcement Ratios. ACI Special
Publication, 2007. 246.
8. Vollum, R., Multipliers for deflections in reinforced concrete flat slabs. Magazine of Concrete
Research, 2003. 55(2): p. 95-104.
9. Rahman, M., T. Ayano, and K. Sakata, The use of physical phenomena to predict time effects in
cracked reinforced concrete beams. Magazine of Concrete Research, 1998. 50(3): p. 219-227.
10. Neville, A., W. Houghton-Evans, and C. Clarke, Deflection control by span/depth ratio.
Magazine of Concrete Research, 1977. 29(98): p. 31-41.
11. Bate, S.C.C., Handbook on the Unified Code for structural concrete (CP11O: 1972). 1972:
Cement and Concrete Association.
12. Scanlon, A. and P.H. Bischoff, Shrinkage restraint and loading history effects on deflections of
flexural members. ACI Structural Journal, 2008. 105(4).
13. Xu, Y., et al., Mechanical properties of expanded polystyrene lightweight aggregate concrete
and brick. Construction and Building Materials, 2012. 27(1): p. 32-38.
14. Trussoni, M., C.D. Hays, and R.F. Zollo, Fracture Properties of Concrete Containing Expanded
Polystyrene Aggregate Replacement. ACI Materials Journal, 2013. 110(5).
15. Sabaa, B. and R.S. Ravindrarajah. Engineering properties of lightweight concrete containing
crushed expanded polystyrene waste. in Materials Research Society, 1997, Fall Meeting,
Symposium MM, Advances in Materials for Cementitious Composites December. 1997.
16. Li, Y., N. Liu, and B. Chen, Properties of lightweight concrete composed of magnesia phosphate
cement and expanded polystyrene aggregates. Materials and Structures, 2015. 48(1-2): p. 269-
276.
17. Gilbert, R. and X. Guo, Time-dependent deflection and deformation of reinforced concrete flat
slabs-an experimental study. ACI structural journal, 2005. 102(3).
18. Ian Gilbert, R., Tension Stiffening in Lightly Reinforced Concrete Slabs. Journal of Structural
Engineering, 2007. 133(6): p. 899-903.
19. AS-3600-09, Concrete structures. Australian Standard, 2009.
20. CSA.A23.3-04, Design of concrete structures. Canadian Standards Association (CSA), 2010.
21. Aslani, F., Experimental and numerical study of time-dependent behaviour of reinforced self-
compacting concrete slabs. PhD dissertation, University of Technology Sydney (UTS), 2014.

383
Chapter 9 – Flexural long-term deflection

22. Nejadi, S., Time-dependent cracking and crack control in reinforced concrete structures. 2005,
PhD thesis, The University of New South Wales Sydney, Australia.
23. 1012.14, A., Methods of testing concrete - Method for securing and testing cores from hardened
concrete for compressive strength. Standards Australia, 1991.
24. C39, A., Standard Test Method for Compressive Strength of Cylindrical Concrete Specimens.
american Society of Testing and Materials, 2000.
25. 1012.17-97, A., Methods of testing concrete - Determination of the static chord modulus of
elasticity and Poisson's ratio of concrete specimens. Standards Australia, 2014.
26. 14, A.C.C.M.-. Standard Test Method for Static Modulus of Elasticity and Poisson’s Ratio of
Concrete in Compression. American Society of Testing and Materials, 2000.
27. 318-08, A., Building code requirements for structural concrete (ACI 318-11) and commentary.
ACI Committee, Standard ACI 318. American Concrete Institute, Farmington Hills, Mich, 2008.
28. Ghali, A., Deflection of reinforced concrete members: a critical review. ACI Structural Journal,
1993. 90(4): p. 364-373.

384
Chapter 10- Numerical analysis

CHAPTER 10

FINITE ELEMENT MODELLING OF


INSTANTANEOUS AND LONG-TERM
DEFLECTION OF LIGHTWEIGHT CONCRETE
SLABS
Chapter 10- Numerical analysis

Table of Contents

10.1. INTRODUCTION .................................................................................................................. 385


10.2. FINITE ELEMENT MODELS FOR REINFORCED CONCRETE STRUCTURES ............ 387
10.3. DAMAGE MODELLING IN CONCRETE ........................................................................... 387
10.3.1. Smeared crack concrete model ........................................................................................... 388
10.3.1.1. Tension before cracking ............................................................................................ 389
10.3.1.2. Tension after cracking ............................................................................................... 390
10.3.2. Damage plasticity method: ................................................................................................. 392
10.4. REINFORCEMENT MODELLING ...................................................................................... 393
10.5. EFFECT OF ANALYSIS STEPS........................................................................................... 394
10.6. CONVERGENCE METHODS IN NON-LINERA ANALYSIS ........................................... 395
10.7. ATENA SOFTWARE ............................................................................................................ 396
10.8. TECHNICAL SPECIFICATIONS OF ATENA-V5 .............................................................. 397
10.8.1. Materials modelling ............................................................................................................ 397
10.8.2. Loading models .................................................................................................................. 398
10.8.3. Finite Elements ................................................................................................................... 398
10.8.4. Solution methods ................................................................................................................ 398
10.8.5. Graphical user environment ............................................................................................... 399
10.9. CONCRETE CONSTITUTIVE MODELS IN ATENA......................................................... 399
10.9.1. Concrete Stress-Strain Relations ........................................................................................ 399
10.9.2. Compressive Strength of Cracked Concrete ....................................................................... 400
10.9.3. Bond-slip models in ATENA ............................................................................................. 401
10.9.4. Crack Models in ATENA ................................................................................................... 402
10.10. MODELLING OF REINFOCEMENT IN ATENA ............................................................... 404
10.11. CRITERIA TO SELECT PROPER ELEMENTS IN ATENA ............................................... 404
10.11.1. Plane elements ............................................................................................................... 405
10.11.1.1. Truss 2D and 3D Element ..................................................................................... 405
10.11.1.2. Plane Quadrilateral Elements ................................................................................ 405
10.11.1.3. Plane Triangular Elements .................................................................................... 406
10.11.2. 3D Solid Elements ......................................................................................................... 406
10.11.2.1. Tetrahedral elements ............................................................................................. 406
10.11.2.2. Brick elements ....................................................................................................... 406
10.11.2.3. Wedge elements .................................................................................................... 407
10.11.3. Spring Element............................................................................................................... 407
10.12. CREEP AND SHRINKAGE ANALYSIS .............................................................................. 411
10.13. MODELIING OF SHRINKAGE BEHAVIOUR IN ATENA ................................................ 413
10.14. SELECTION OF COMPATIBLE ELEMENTS WITH EXPERIMENTAL DATA.............. 414
10.15. EFFECT OF MAXIMUM AGGREGATE SIZE AND AGGREGATE INTERLOCK ......... 415
10.16. EFFECT OF MESH SIZE ...................................................................................................... 416
10.17. WIDTH AND HEIGHT OF THE SELECTED ELEMENTS ................................................ 417
Chapter 10- Numerical analysis

10.18. EFFECT OF LOADING PADS IN PROGRAMMING ......................................................... 417


10.19. ADVANTAGES OF SLABS SYMMETRY IN NUMERICAL ANALYSIS ........................ 418
10.20. ANALYSIS OF SLABS UNDER INSTANTANEOUS LOADING ..................................... 419
10.20.1. Process of the modelling ................................................................................................ 419
10.20.2. Slab dimensions ............................................................................................................. 420
10.20.3. Basic material properties ................................................................................................ 420
- CC3DCementitious material: .................................................................................................. 421
- CC3DNonLinCementitious material: ..................................................................................... 421
- CC3DNonLinCementitious2 material: ................................................................................... 421
- CC3DNonLinCementitious2Variable material: ...................................................................... 421
- CC3DNonLinCementitious2User material: ............................................................................ 421
10.20.4. Solution parameters in instantaneous loading analysis .................................................. 424
10.20.5. Monitoring points of instantaneous deflection ............................................................... 425
10.20.6. Effect of mesh size on deflection and cracking of slabs ................................................ 426
10.20.7. Effect of bond-slip model .............................................................................................. 427
10.20.8. Ratio of deflection at monitoring points ........................................................................ 429
10.20.9. Effect of support condition on the instantaneous deflection .......................................... 430
10.20.10. Sensitivity of deflection to early age properties of concrete .......................................... 430
10.20.11. Effect of loading pads .................................................................................................... 431
10.21. TIME-DEPENDENT ANALYSIS OF SLABS ...................................................................... 433
10.21.1. Modelling of the slab ..................................................................................................... 433
10.21.2. Basic material properties ................................................................................................ 434
10.21.3. Slab dimensions ............................................................................................................. 435
10.21.4. Monitoring points for long-term deflection ................................................................... 436
10.21.5. Loading steps ................................................................................................................. 436
10.21.6. Verification of the proposed models .............................................................................. 438
10.21.7. Plane stress and plane strain analysis ............................................................................. 440
10.21.7.1. Effect of depth of structure in 2D analysis ............................................................ 441
10.21.8. Comparing plane stress, plane strain and 3D elements .................................................. 446
10.22. EFFECT OF THE IMPLEMENTED CREEP MODELS ON THE LONG-TERM
DEFLECTION OF SLABS ...................................................................................................................... 447
10.22.1. B3 creep model of ATENA ........................................................................................... 448
10.22.2. FIBMC-2010 creep model of ATENA........................................................................... 450
10.22.3. BP-KX creep model of ATENA .................................................................................... 451
10.22.4. CEB-FIP78 creep model of ATENA ............................................................................. 452
10.22.5. CSN-731201 creep model of ATENA ........................................................................... 453
10.22.6. BP1 creep model of ATENA ......................................................................................... 455
10.22.7. BP2 creep model of ATENA ......................................................................................... 456
10.22.8. Comparing the creep models of ATENA ....................................................................... 457
10.23. SUMMARY ....................................................................................................................... 459
REFERENCES ......................................................................................................................................... 462
Chapter 10- Numerical analysis

List of Figures

Figure 10.1: Uniaxial stress-strain response and cracking of concrete [2] ............................................... 389
Figure 10.2: Biaxial Failure criteria of concrete [9] ................................................................................. 390
Figure 10.3: Exponential crack opening model of concrete in ATENA [9] ............................................. 391
Figure 10.4: Typical tension stiffening model in FEA simulation [2] ..................................................... 392
Figure 10.5: Different methods of reinforcing modelling in FEM [11] ................................................... 393
Figure 10.6: Interaction law for reinforcing steel in FEA ........................................................................ 394
Figure 10.7: Compressive strength reduction of cracked concrete [5] ..................................................... 400
Figure 10.8: Reinforcement and bond details in ATENA (2014) 2D analysis [5] ................................... 402
Figure 10.9: Reinforcement and bond details in ATENA (2014) 3D analysis [5] ................................... 403
Figure 10.10: Smeared crack models in ATENA [5] ............................................................................... 403
Figure 10.11: Geometry of CCIsoTruss elements [5] .............................................................................. 405
Figure 10.12: Geometry of CCIsoQuad elements (by 4 to 9 nodes) [5] ................................................... 405
Figure 10.13: Geometry of CCIsoTriangle elements (by 3 to 6 nodes) [5] .............................................. 406
Figure 10.14: Geometry of CCIsoTetra elements (by 4 to 10 nodes) [5] ................................................. 406
Figure 10.15: Geometry of CCIsoBrick elements (by 8 to 20 nodes) [5] ................................................ 407
Figure 10.16: Geometry of CCIsoWedge elements (by 8 to 20 nodes) [5] .............................................. 407
Figure 10.17: Geometry of a) 2D CCSpring and CCIsoLineSpring and b) 3D CCSpring and
CCPlaneSpring [5] ................................................................................................................................... 408
Figure 10.18: Shrinkage input data implementation in ATENA-CREEP analysis ................................. 414
Figure 10.19: Implementation of aggregate interlock and maximum size of aggregate in FEA in ATENA
.................................................................................................................................................................. 416
Figure 10.20: Simulated and real loading pads ........................................................................................ 418
Figure 10.21: Half-slab model of slab in GID environment ..................................................................... 418
Figure 10.22: Details, geometry and meshing of the LWC slabs in GID program .................................. 420
Figure 10.23: Material properties for CC3DNonLinCementitious2 ......................................................... 422
Figure 10.24: Parameters to define the reinforcement in GID-ATENA................................................... 423
Figure 10.25: Typical convergence of the model analysis in ATENA program ...................................... 424
Figure 10.26: Crack propagation and deflection in the mid third area of the slab span by different mesh
sizes .......................................................................................................................................................... 427
Figure 10.27: Cracking pattern and propagation in slab spans by different mesh sizes ........................... 427
Figure 10.28: Ratio of deflection by different mesh sizes to deflection by 7 mm mesh .......................... 428
Figure 10.29: Ratio of the maximum crack with at different mesh sizes to crack with by 7 mm mesh ... 429
Figure 10.30: Localized stress and strain distribution around the loading plate (flexible plates) ............ 431
Figure 9.31: Localized stress and strain distribution around the loading plate (Rigid plates) .................. 432
Figure 10.32: Different problem types in ATENA to analyse the model ................................................. 433
Figure 10.33: Implemented creep models in ATENA-creep problem type .............................................. 434
Figure 10.34: Input data table to define the interval data in creep analysis ............................................. 437
Figure 10.35: Typical deflection-time diagram of LWC slab in ATENA environment ........................... 437
Figure 10.36: comparison of the proposed model with the experimental data in loading level 1 ............ 438
Figure 10.37: Comparison of the proposed model with experimental data in loading level 2 ................. 439
Figure 10.38: Plane stress (a) and plane strain (b) presentation of a general element in the structure ..... 440
Figure 10.39: Typical input data table for creep analysis in GID-ATENA .............................................. 442
Figure 10.40: Deflection versus time in the model with different thickness values of plane strain element
.................................................................................................................................................................. 443
Figure 10.41: Deflection versus time in the model with different thickness values of plane stress element
.................................................................................................................................................................. 443
Figure 10.42: Ratio of deflection of plane stress element models to the experimental deflection at 500
days .......................................................................................................................................................... 444
Figure 10.43: Deflection versus time in the model with different thickness values of plane stress element
.................................................................................................................................................................. 445
Chapter 10- Numerical analysis

Figure 10.44: Ratio of deflection of plane stress element models to the experimental deflection at 500
days .......................................................................................................................................................... 445
Figure 10.45: Effect of loading level and thickness of plane stress element on maximum deflection ..... 446
Figure 10.46: Comparing the plane stress, plane strain and 3D elements ................................................ 447
Figure 10.47: Options to activate the shrinkage and compliance effects in creep analysis in ATENA ... 448
Figure 10.48: B3 creep model with and without shrinkage and compliance effects in loading level 1.... 449
Figure 10.49: B3 creep model with and without shrinkage and compliance effects in loading level 2.... 449
Figure 10.50: FIBMC-2010 creep model with and without shrinkage and compliance effects in loading
level 1 ....................................................................................................................................................... 450
Figure 10.51: FIBMC-2010 creep model with and without shrinkage and compliance effects in loading
level 2 ....................................................................................................................................................... 451
Figure 10.52: BP-KX creep model with and without shrinkage and compliance effects in loading level 1
.................................................................................................................................................................. 452
Figure 10.53: BP-KX creep model with and without shrinkage and compliance effects in loading level 2
.................................................................................................................................................................. 452
Figure 10.54: CEB-FIP78 creep model with and without shrinkage effect in loading level 1 ................. 453
Figure 10.55: CEB-FIP78 creep model with and without shrinkage effect in loading level 2 ................. 453
Figure 10.56: CSN-731201 creep model with and without shrinkage effect in loading level 1 ............... 454
Figure 10.57: CSN-731201 creep model with and without shrinkage effect in loading level 2 ............... 454
Figure 10.58: BP1creep model with and without shrinkage effect in loading level 1 .............................. 455
Figure 10.59: BP1creep model with and without shrinkage effect in loading level 2 .............................. 455
Figure 10.60: BP2 creep model with and without shrinkage effect in loading level 1 ............................. 456
Figure 10.61: BP2 creep model with and without shrinkage effect in loading level 2 ............................. 457
Figure 10.62: Comparing the implemented creep models in loading level 1 ........................................... 458
Figure 10.63: Comparing the implemented creep models in loading level 2 ........................................... 458
Figure 10.64: comparison of the proposed model with the experimental data in loading level 1 ............ 460
Figure 10.65: Comparison of the proposed model with experimental data in loading level 2 ................. 461
Chapter 10- Numerical analysis

List of Tables
Table 10.1: Available element types in ATENA [5] ................................................................................ 409
Table 10.2: Material types in ATENA [5]................................................................................................ 411
Table 9.3: Material models for creep in TANEA [5] ............................................................................... 412
Table 10.4: Element type and material compatibility [5] ......................................................................... 413
Table 10.5: Mechanical properties of lightweight concrete ..................................................................... 420
Table 10.6: Mechanical properties of steel bars ....................................................................................... 421
Table 10.7: Variable parameters in definition of typical concrete properties in GID-ATENA [5] .......... 423
Table 10.8: Solution parameters in instantaneous deflection analysis of LWC slabs .............................. 424
Table 10.9: Experimental instantaneous deflection values under LL1 and LL2 in monitoring points ..... 425
Table 10.10: Variable arrangement of parameters in SBETA material and the obtained deflections in LL1
.................................................................................................................................................................. 426
Table 10.11: Preferences in studying the effect of mesh size on instantaneous deflection ...................... 426
Table 10.12: Ratio of deflection in corresponding points of simply-supported and fixed ends slab ........ 430
Table 10.13: deflection ratio at the corresponding point by concrete properties at 14 and 28 days ......... 431
Table 10.14: Input parameters in the creep and shrinkage models implemented in ATENA .................. 435
Table 10.15: Stress and strain assumptions in plane stress and plane strain conditions ........................... 441
Chapter 10- Numerical analysis

Chapter 10: Numerical analysis

10.1. INTRODUCTION

To develop advanced design and analysis methods, the need for experimental research is
crucial. Experimental results provide a secure basis for verification of proposed design
equations. Results of the empirical study also supply the required basic data and information for
Finite Element (FE) models and programming for analysing the structure. Furthermore, the
results of FE models have to be utilized to assess the structure and proposed models should be
calibrated with experimental results.

Since the first FE study in reinforced concrete members conducted by Ngo and Scordelis
(1967) [1], analyses of reinforced concrete structures using Finite Element Method (FEM), have
experienced a remarkable improvement. Development of appropriate analytical models can
decrease the number of required empirical activities carried out upon the test specimens or even
the real structures for the solution of a given problem, also; identifying that these tasks are time-
consuming, very expensive and often do not simulate exactly the loading and support conditions
of the actual structure.

FEM has thus become a powerful computational tool, allowing complex analyses of the
nonlinear response of Reinforced Concrete (RC) structures to be carried out in a routine fashion.
However, the complex behaviour of concrete sets some limitations in implementing FEM. The
complexity is mainly due to[2]:

- Non-linear stress-strain relation of the concrete under multi-axial stress conditions;


- Strain softening and anisotropic stiffness reduction;
- Progressive cracking caused by tensile stresses and strains;

385
Chapter 10- Numerical analysis

- Bond between concrete and reinforcement;


- Aggregate interlocks and dowel action of reinforcement;
- Time-dependent behaviour as creep and shrinkage.

The nonlinear Finite Element Analysis (FEA) can define crack formation and growth, the
appearance and process of structure breakage, and additionally assess the ultimate bearing
capacity and reliability of the structure, which is important in analysis and design of the
structure [3].

Constitutive relation of the reinforced concrete material has considerable influence on the
nonlinear analysis. An appropriate material model in the FE model should certainly be able to
represent both elastic and plastic behaviour of concrete in compression and tension.

Chen and Saleeb (1982); Parton and El-Barbary (1983); Vidosa et al. (1988); Marzouk and
Chen (1993); Marzouk and Jiang (1996); Al-Nasra (1997); Jiang and Mirza (1997); Reitman
and Yankelevsky (1997); Polak (1998); Huang et al. (1999); Enochsson and Dufvenberg
(2001); Shanmugam et al. (2002); Vainiunas et al. (2004); Polak (2005); Murray et al. (2005)
and Deaton (2005) are just some researchers who have applied FE models to investigate the
interaction of different non-linear effects on the response of reinforced concrete structures[4].

The aim of this study is to simulate and compare the effective modelling of reinforced
concrete one-way slabs made with Light-Weight Concrete (LWC) containing treated Expanded
Poly-Styrene (EPS) beads. FEM has been used to model the slabs with the aid of a computer
program to process the FE model. The models were assessed on their capability of replicating
the real cracking and deflection behaviour in the real models properly. There are multiple
methods for determining the deflection of beams as well as calculating crack spacing and crack
width concerning the instantaneous (short-term) loading or after being subjected to the sustained
load for a long time. This study utilizes the FE based computer program, ATENA, developed by
Cervenka et al., (2015) [5] to simulate the reinforced concrete slabs. Mechanical properties of
concrete including the time-dependent properties have been simulated by the ATENA program
also. The main objectives of this chapter are:

- A brief presentation and discussion of the basic concepts, materials modelling and
fracture models in general FEMs and particularly, ATENA;
- Non-linear FE modelling of the reinforced concrete structures subjected to sustained
service loads;
- Parametric study of the hardened concrete properties models, bond-slip models, and
creep and shrinkage models based on the experimental results and the default ATENA
models, to simulate and determination of instantaneous and time-dependent deflection of
lightweight concrete slabs.
386
Chapter 10- Numerical analysis

In addition, the numerical analysis and parametric study of the slabs in this research contains
the following items:
1- Comparing the instantaneous deflection of LWC slabs under two different levels of
loading;
2- Comparing the time-dependent deflection of LWC slabs under the same two distinct
levels of loading as for the instantaneous deflection;
3- Sensitivity of the instantaneous and time-dependent deflection to the implemented
material models in ATENA;
4- Investigating the ratio of time-dependent deflection to the instantaneous deflection under
two different levels of loading and comparing these with the experimental results to
verify the corresponding multipliers in codes of practice.

10.2. FINITE ELEMENT MODELS FOR REINFORCED CONCRETE


STRUCTURES

Constitutive concrete models are implemented in several software applications. These


analysis programs contain possible approaches to the behaviour of concrete elements and
concrete models. To choose the proper model of concrete, considering the purpose and the area
where the concrete should be used is mandatory; because the concrete models of different
purposes require different data and accuracies of the models. The required details of the
computational model are an example of this matter. In the majority of the advanced models, the
more accurate the model, the higher the demands regarding the extent and quality of the
complex data, which makes, in turn, the computational process more demanding. A non-linear
analysis of reinforced and plain concrete is often combined with other types of the task. Such
analysis tasks could, for instance, simulate dynamic or earthquakes loads.

Developing a constitutive model which can provide reliable predictions of short-term and
long-term deflection behaviour of reinforced LWC to facilitate a parametric study of deflection
behaviour is the primary scope of this study. In this regard, the effective parameters of the
deflection variation with time may be identified and quantified.

10.3. DAMAGE MODELLING IN CONCRETE

There are three different models used to simulate the damage in reinforced concrete elements:
(1) Smeared crack concrete model, (2) Brittle crack concrete model, and (3) Concrete damaged
387
Chapter 10- Numerical analysis

plasticity model [6]. Smeared crack and damage plasticity models are two main methods of the
non-linear analysis of concrete in FE simulations.

10.3.1. Smeared crack concrete model

In this model, the cracking process begins at any location where the concrete stresses reach
one of the failure surfaces either in the biaxial tension region or a combined tension-
compression region and it does not track individual “macro” cracks. Developed cracks affect the
stress and material stiffness associated with the integration point. Cracking is assumed to be the
most important aspect of the concrete behaviour and the representation of cracking and post-
cracking anisotropic behaviour dominates the modelling [2]. This model is proper for the
concrete which is subjected to essentially monotonic straining at low-confining pressures. It
uses oriented damaged elasticity concepts (smeared cracking) to describe the reversible part of
the response of material after cracking failure [7].

The strain is decomposed into elastic strain and plastic strain as illustrated in Equation (10.1).

ߝ=ߝ݁+ߝ‫݌‬ Eq. (10.1)

Where; ߝ݁=ߪ/‫ ܿܧ‬is elastic strain, and ߝ‫ ݌‬is plastic strain.

The uniaxial compressive stress-strain relationship proposed by Saenz (1964) is now widely
accepted as the mathematical description of the uniaxial stress-strain curve for concrete [8]. The
cracking and compressive responses included in the concrete model are demonstrated by the
uniaxial response of a specimen as shown in Figure (10.1). When concrete is subjected to
compressive stresses, the non-linear deformations of concrete are basically inelastic, because
upon unloading, only a portion of those strains can be recovered from the total strains.
Therefore, the stress-strain behaviour of the concrete material may be separated into recoverable
and non-recoverable components. The recoverable part can be treated within the field of
elasticity theory while the theory of plasticity can treat the irrecoverable part. Plasticity-based
models have been used extensively in recent years to describe the behaviour of concrete [8].

388
Chapter 10- Numerical analysis

Figure 10.1: Uniaxial stress-strain response and cracking of concrete [2]

A “compression” failure surface forms the basis of the model for the non-linear response
when the principal stresses are mainly compressive. When the stress reaches a failure surface
which is called the “crack detection” surface in tension (including tension-compression zone),
cracking is expected to happen.

10.3.1.1. Tension before cracking

The behaviour of concrete in tension without cracks is assumed to be linear elastic. The
equivalent uniaxial law is used to describe the non-linear behaviour of concrete in the biaxial
state regarding the equivalent uniaxial strain, εeq and the effective stress σcef in Equation (10.2).

௘௙ ᇱ௘௙
ߪ௖ ൌ ‫ܧ‬௖ ߝ ௘௤ Ǣ Ͳ ൑ ߪ௖ ൑ ݂௧ ‫ݍܧ‬Ǥ ሺͳͲǤʹሻ

where; ˆ୲ᇱୣ୤ is the effective tensile strength derived from the biaxial failure function of
concrete and Ec is the initial elastic modulus of concrete.

As shown in Figure (10.2), the biaxial failure of concrete contains distinct tension and
compression failure criteria. In the compression-compression stress state, the failure is a
function of the principal stresses in concrete (σc1, σc2) and the uniaxial cylinder strength of
concrete (ˆୡᇱ ). In the tension-compression state, the failure function continues linearly from the
point σc1 = 0, σc2 =ˆୡᇱ into the tension-compression region with the linearly decreasing strength as
described in Equation (10.3).

389
Chapter 10- Numerical analysis

ᇱ௘௙
݂௖ ൌ ‫ݎ‬௘௖ Ǥ ˆୡᇱ Ǣ ͲǤͻ ൑  ‫ݎ‬௘௖ ൑ ͳǤͲ‫ݍܧ‬Ǥ ሺͳͲǤ͵ሻ

Where; rec the reduction factor of the compressive strength is a function of σc2 and ˆୡᇱ .

Figure 10.2: Biaxial Failure criteria of concrete [9]

In the tension-tension area, the tensile strength is constant and equal to the uniaxial tensile
strength (݂୲ᇱ ). In the tension-compression state, the tensile strength is reduced as described in
Equation (10.4).

ᇱ௘௙
݂௧ ൌ ‫ݎ‬௘௧ Ǥ ˆ୲ᇱ ‫ݍܧ‬Ǥ ሺͳͲǤͶሻ

Where; ret the reduction factor of the tensile strength is a function of σc2 and ˆୡᇱ .

10.3.1.2. Tension after cracking

After cracking, a model based on a crack opening law and fracture energy is used in
combination with the crack band theory. Gilbert et al. (2010) [9] used the exponential model in
Equation (10.5) to describe the post-cracking behaviour of concrete. Other existing models in
literature have been discussed in Chapter 4 of this study.

390
Chapter 10- Numerical analysis

ߪ ‫ ݓ‬ଷ ି௖మ ௪௪ ‫ݓ‬


ᇱ௘௙
ൌ ቊͳ ൅ ൬ܿଵ ൰ ቋ݁ ೎ െ ሺͳ ൅ ܿଵଷ ሻ ݁ ି௖మ Ǣ
݂௧ ‫ݓ‬௖ ‫ݓ‬௖
ீ೑
ܿଵ ൌ ͵Ǣܿଶ ൌ ͸Ǥͻ͵Ǣ‫ݓ‬௖ ൌ ͷǤͳͶ ᇲ೐೑ ‫ݍܧ‬Ǥ ሺͳͲǤͷሻ
௙೟

Where; ‫ ݓ‬is the crack width; ‫ݓ‬௖ is the crack width at the complete release of stress; ߪ is the
normal stress on the crack (crack cohesion); ‫ܩ‬௙ is the fracture energy needed to create a unit
ᇱ௘௙
area of stress-free crack and was assigned a value of 25 N/m in the numerical model; and ݂௧ t
is the effective tensile strength derived from the failure function.

The crack opening displacement ‫ ݓ‬is derived from strains according to the crack band theory.
Figure (10.3) shows the exponential crack opening law in concrete. This law is just one of the
five alternatives available in the utilized program, ATENA.

Figure 10.3: Exponential crack opening model of concrete in ATENA [9]

Once the crack occurs, the FE program uses a smeared crack approach in which constitutive
calculations are performed independently at each integration point of the FE model. The
initiation of cracks enters into these calculations by affecting the stress and material stiffness
associated with integration points [2].

The smeared crack concrete model applies tension stiffening to specify the post-failure stress-
strain relationship. Tension stiffening also can be modelled by applying a fractured energy
cracking criterion. A typical tension stiffening model is shown in Figure (10.4).

391
Chapter 10- Numerical analysis

Figure 10.4: Typical tension stiffening model in FEA simulation [2]

Proper application of tension stiffening parameters is essential, as tension stiffening values


which are too small cause the local cracking failure in the concrete to introduce temporarily
unstable behaviour in the overall response of the model while high tension stiffening makes it
easier to obtain numerical solutions.

10.3.2. Damage plasticity method:

The damage in quasi-brittle materials in which the confining pressure is sufficiently large to
prevent the brittle fracture can be defined by evaluating the dissipated fracture energy required
to create micro cracks [2]. Strain rate decomposition is assumed for the rate-independent model
as presented in Equation (10.1).

The stress-strain relations are governed by scalar damaged elasticity in Equation (10.6):

ɐ ൌ ሺͳ െ Ɍሻୣ୪ ୣ୪
଴  ‫  ׷‬൫ɂ െ ɂ௣ ൯ ൌ   ‫  ׷‬൫ɂ െ ɂ௣ ൯‫ݍܧ‬Ǥ ሺͳͲǤ͸ሻ

Where; ‫ܦ‬଴௘௟ is the initial (undamaged) elastic stiffness of the material; ߪ is the degraded
elastic stiffness; ξ is the scalar stiffness degradation variable, which can take variation in the
range from zero (undamaged material) to one (fully damaged material).

The associated damage with the failure mechanisms of the concrete (cracking and crushing)
results in a reduction in the elastic stiffness. Within the context of the scalar-damage theory, the
stiffness degradation is isotropic and described by a single degradation variable, d. Following

392
Chapter 10- Numerical analysis

the usual concepts of continuum damage mechanics, the effective stress is defined as Equation
(10.7):

ɐ௘ ‫  ؝‬ୣ୪
଴  ‫  ׷‬ሺɂ െ ɂ௣ ሻ Eq. (10.7)

The Cauchy stress is associated with the effective stress through the scalar degradation
relationship as explained in Equation (10.8):

ɐ ൌ ሺͳ െ Ɍሻɐ௘ Eq. (10.8)

10.4. REINFORCEMENT MODELLING

There are three approaches to how to model the steel bars in FE modelling; distributed steel
approach, embedded steel approach and discrete steel approach. The general view of these
models is presented in Figure (10.5). The classic model [10] is generally used for the elastic and
plastic behaviour of concrete under tension and compression. A multi-linear stress-strain
relationship is used to model the reinforcement in the numerical model and allows all stages of
steel behaviour to be readily modelled, including the elastic state, yield plateau, strain
hardening, and fracture [9, 10].

Distributed steel approach Embedded steel approach Discrete steel approach

Figure 10.5: Different methods of reinforcing modelling in FEM [11]

Figure (10.6) shows the typical bilinear and multi-linear stress-strain relationship of steel bars
in FEM.

393
Chapter 10- Numerical analysis

Bilinear load-deflection [10] Multi-linear stress-strain [5]


Figure 10.6: Interaction law for reinforcing steel in FEA

10.5. EFFECT OF ANALYSIS STEPS

Finite element simulations enable designers and researchers to consider problems that exhibit
some form of non-linear behaviour. Whether it is large-scale deformation, plasticity, or contact
between bodies, we generally classify these non-linear problems as containing some form of
varying stiffness in the model. Sometimes, this change in flexibility can be abrupt-like bodies
coming into contact with each other, or smooth-like creep or low levels of plasticity. With all of
these problems, a representative force-deflection curve is not linear and the stiffness changes
with deformation.

In general, geometric non-linearity, material non-linearity, force-boundary condition


nonlinearity, and displacement boundary condition non-linearity are the common sources of
non-linearity. Non-linear FEA applies to the following conditions:

x Stability analysis: To find critical points (limit points and bifurcation points) closest
to operational range;
x Service configuration analysis: To find the ‘operational’ equilibrium configuration of
certain slender structures when the fabrication and service configurations are quite
different (e.g. cable and inflatable structures);
x Reserve strength analysis: To find the load-carrying capacity beyond critical points to
assess safety under abnormal conditions;
x Progressive failure analysis: A combined strength and stability analysis in which
progressive deterioration (e.g. cracking) is considered.

For static FEA, the equation representing equilibrium can be written in the matrix form of
Equation (10.9):
394
Chapter 10- Numerical analysis

‫ܭ‬Ǥ οܷ ൌ ܰ െ ‫ݍܧܨ‬Ǥ ሺͳͲǤͻሻ

Where N represents the vector of applied loads, F is the vector of internal forces, and ΔU is
the vector of current nodal displacement. In non-linear analysis, the load N is used in a series of
load steps N1, N2, N3,….Nn.

Finite Element Analysis (FEA) involves solving the Equation (10.9) for AU. For the n-th load
step, the equation is often solved by iterations of the form presented in Equation (10.10)

‫ܭ‬Ǥ οܷሺ௜ାଵሻ ൌ ܰሺ௡ሻ െ ‫ܨ‬ሺ௜ሻ ‫ݍܧ‬Ǥ ሺͳͲǤͳͲሻ

There are different methods of solving the non-linear equations in FEA. By increasing the
load value in higher load steps until the cracking load, the concrete structures often show a
plastic flow and even softening behaviour, while deformation is increased quickly with slight
load change. Since the load is not sensitive, if the force is used as a convergent norm, it is often
difficult to achieve convergence. So the displacement is used as the convergent norm during the
loading process [12].

FE programs have taken on the task of solving non-linear static structural problems in a
number of ways. In all cases, the algorithms have to iterate to achieve equilibrium for static
structural problems. The solver will apply some portion of the external forces to the model, and
then calculate the resultant internal forces to balance the applied load. There will initially be
some out of balance residual force so the solver will have to improve the stiffness to narrow
down the imbalance difference toward zero. Obtaining final convergence and equilibrium, when
running these iterative problems, can be very challenging.

10.6. CONVERGENCE METHODS IN NON-LINERA ANALYSIS

A convergence criterion measures how well the obtained solution satisfies the non-linear
equilibrium. In FEA, the convergence criteria are usually based on some norms of the
displacement, residuals and energy as the product of residual and displacement.

The convergence criteria and tolerances must be carefully chosen so as to provide accurate
yet economical solutions. If the convergence criterion is too broad, inaccurate results are
obtained. If the convergence criterion is too tight, too much effort will be spent in obtaining
unnecessary accuracy.
395
Chapter 10- Numerical analysis

10.7. ATENA SOFTWARE

Among the existing FE software examining the accurate simulation of reinforced concrete
structures, this study utilizes the ATENA program for analysis of the short-term and long-term
behaviour of the reinforced LWC one-way slabs.

Application of ATENA for instantaneous, time-dependent and fracture analysis of reinforced


concrete structures has been increasing, especially in recent years. Cervenka et al. (2002) [13],
Johnson (2006) [14], Yang et al. (2007) [15], Yu (2007) [16], Malm and Holmgren (2008)[17],
Talley (2009)[18], GribniaK et al. (2010) [19], Gilbert and Sakka (2010)[19], Dobromil et al.
(2010)[20], Olteanu et al. (2011) [21], Lampropoulos et al. (2012) [22], Sucharda and
Brozovsky (2014) [23], Singh ET AL. (2014) [24], Meskenas et al. (2014)[25] and Namy
(2015) [26], are just some researchers who have used ATENA in their investigations.

The main reasons for choosing ATENA in the current study can be justified as follows:

a) Almost all of the creep models in the widely used codes of practice and international
standards are included in ATENA and can be used during the analysis;

b) By utilizing the ATENA program it is possible to simulate the shrinkage behaviour of


reinforced concrete members precisely;

c) In addition to the available models for different properties of concrete, there is a facility to
define extra properties or modify the available models by the user. Hardened mechanical
properties of concrete such as compressive strength, tensile strength, modulus of elasticity,
stress-strain relationship, bond stress, tension stiffing factor and cark width are some user-
defined parameters in this program;

d) A wide range of structural, non-structural, fibre reinforced and special concrete can be used
by this program;

e) The ATENA program package includes ATENA-GiD and the ATENA-FE program. The
main ATENA-FE program is used for analysis of the simulated model, while the software GiD
is parallel to ATENA and can be utilized for interfacing with ATENA for data preparation and
mesh generation.

In comparison to the multiple choices of material models, ATENA provides a much more
appropriate selection of models for simulating requirements. Moreover, in most cases, ATENA
also offers the option of using either a rotating crack model or a fixed crack model [7].

Other key features in ATENA are as follows:


ƒ Three-dimensional (3D) environment with User friendly modelling of reinforcement;
ƒ Unique visualization of crack propagation;
396
Chapter 10- Numerical analysis

ƒ Real-time display of results even during the non-linear analysis based on FE method
and fracture mechanics;
ƒ Advanced material models for concrete, reinforcement, steel, rock, soil and masonry;
ƒ Support for the analysis of modern fibre reinforced concrete materials: Strain Hardening
Cementitious Composites (SHCC), Engineered Cementitious Composite (ECC), High-
Performance Fibre Reinforced Concrete (HPRFC), Ultra High-Performance Fibre
Reinforced Concrete (UHPFRC);
ƒ Dynamics, statics, creep, thermal and moisture analyses.

10.8. TECHNICAL SPECIFICATIONS OF ATENA-V5

10.8.1. Materials modelling

This program offers a wide variety of different material models for various materials and
purposes. The following are general definitions of the existing material models. Detailed
specifications are discussed in next sections:

- Three-dimensional fracture-plastic concrete model based on Menetrey-Willam law: smeared


cracks, fracture-energy based softening, compression field theory, aggregate interlock in shear,
non-associated plasticity, unloading direction, and user defined functions, variable parameters;

- Two-dimensional (2D) “SBETA” concrete model, also for high-strength and Steel Fibre
Reinforced Concrete (SFRC): smeared cracks, crack-band, fracture energy based softening,
Kupfer’s compressive failure, variable shear retention, strength reduction of cracked concrete;

- Reinforcement bi-linear and multi-linear, reinforcement with bond, cyclic reinforcement


(Menegotto-Pinto / Bauschinger effect) and bond;
- Von-Mises plasticity for metals such as steel reinforcement;

- Drucker-Prager plasticity with associated / non-associated flow rule for rock and soil;

- Bazant M4/M7 micro-plane concrete;

- Interface with Mohr-Coulomb material law;

- Isotropic elastic;

- Non-linear springs;

- Temperature-dependent material properties (Fire loading);

- Creep and shrinkage (Bazant B3, Euro code, ACI and others);

- Heat and moisture transport, hydration heat model CERHYD;


397
Chapter 10- Numerical analysis

- Fatigue of concrete in tension;

- User-defined material model (user-compiled DLL in C/C++ or FORTRAN).

10.8.2. Loading models

ATENA enables loading of the structure with various actions. These loading cases are
combined into load steps, which are solved utilizing advanced solution methods like Newton–
Raphson, modified Newton–Raphson or arc-length. The following are the main parts of loading
models in ATENAN:
- Load cases: body forces, nodal or linear forces, supports, prescribed deformations,
temperature, shrinkage, prestressing;
- Load steps: the combination of load cases, with solution methods;
- Arbitrary load history in steps, non-proportional, cyclic, dynamic;
- Construction process.

10.8.3. Finite Elements

ATENA includes a wide variety of finite elements of linear and non-linear analysis of
reinforced concrete structures. The following are the main types of finite element in ATENA:

- 2D iso-parametric elements, quadrilateral, triangular, axisymmetric elements

- 3D solid elements: tetrahedron, brick, wedge; low and high order

- 3D shell elements (layered) and beams (fibre)

- Truss elements for reinforcement, external cable elements

- Spring supports – point, line, surface

- Interface elements, 2D and 3D

10.8.4. Solution methods

There are different methods of solving the non-linear equations in FEA of ATENA program.
The following are the main solution methods in ATENA:

- Direct band (skyline) and sparse iterative equation system solvers; eigenvalues;

- Newton-Raphson, modified Newton-Raphson, arc-length, line-search;

- Tangential and elastic stiffness predictors;


398
Chapter 10- Numerical analysis

- Newmark and Hughes Alpha for dynamics;

- 32-bit and 64-bit solution core.

10.8.5. Graphical user environment

ATENA offers a user-friendly graphical interface, which enables efficient solving of


engineering problems including anchoring technology and reinforcement of concrete:

- 2D Graphical User Environment (GUE) including: pre-processing (geometrical modelling,


reinforcement (bars, smeared), automatic meshing, material properties, loading and supports,
solution methods, monitoring), solution (non-linear FE solution, graphical monitoring, restart),
post-processing (iso-lines, iso-areas, rendering, vectors, tensors, cracks, response diagrams,
cuts/sections, internal forces, user-defined numerical output);

- 3D GUE including: pre-processing (geometrical modelling, reinforcement (bars, smeared),


copy and move, automatic meshing, material properties, loading and supports, solution
methods, monitoring), solution (direct or sparse iterative, graphical monitoring, restart), post-
processing (iso-lines, iso-areas, rendering, vectors, tensors, cracks, response diagrams,
cuts/sections, user-defined numerical output);

- ATENA Studio: Graphical user interface for ATENA Science for solution and post-
processing of 2D and 3D models;

- GiD interface (GiD - general FE pre- and post-processor from CIMNE, Spain) with
interface with ATENA.

10.9. CONCRETE CONSTITUTIVE MODELS IN ATENA

10.9.1. Concrete Stress-Strain Relations

There are different models in ATENA used to simulate the material behaviour of the elastic,
elastic-plastic and plastic range in linear and non-linear based analysis. Also, the time-
dependent variation of materials such as creep and shrinkage behaviour, the elasticity-based
material model (SBETA), the fracture-plastic material model, and creep and shrinkage analysis
algorithms are implemented in ATENA.

The constitutive model of materials used to explain the non-linear behaviour of concrete in a
two-dimensional plane stress state is known as SBETA in ATENA. The fracture of plastic
material as one of the advanced material models is a predecessor program using the SBETA

399
Chapter 10- Numerical analysis

model [23]. This model is founded on the smeared cracks, damage and fracture mechanics
concepts [17].

The material SBETA includes the following effects of concrete behaviour also.

(i) Non-linear behaviour in compression including hardening and softening;


(ii) Fracture of concrete in tension, based on the non-linear fracture mechanics;
(iii) Biaxial strength failure criteria;
(iv) Reduction of compressive strength after cracking;
(v) Tension stiffening effect;
(vi) Cracking models (fixed crack direction and rotated crack direction).

The equivalent uniaxial law is used to describe the non-linear behaviour of concrete in the
biaxial state regarding the equivalent uniaxial strain, and the effective stress [8].

10.9.2. Compressive Strength of Cracked Concrete

The reduced compressive strength of concrete due to cracking can affect the results of
analysis. In ATENA, the reduction of the compressive strength after cracking in the direction
parallel to the cracks is included. This decrease is performed by a Gaussian function as shown in
Figure (10.7). The constant ‘c’ represents the maximal strength reduction under the large
transverse strain.

ᇱ௘௙
In Figure (10.7), ݂௧ is the effective tensile strength derived from the biaxial failure function
of concrete, ݂௖ᇱ is the uniaxial cylinder strength of concrete and rc is the reduction factor.

Figure 10.7: Compressive strength reduction of cracked concrete [5]

400
Chapter 10- Numerical analysis

Reinforcing bars are the main effective parameters on the value of the reduction factor in
reinforced concrete. However, this factor covers the reduction of compressive strength in
bridges in plain concrete; when the strain localizes in one main crack, the compressive concrete
struts can cross this crack (bridging effect).

Kolleger et al. (1988) [27] proposed rc equal to 0.45 for reinforced concrete with fine mesh.
In the investigation by Dyngeland (1989)[28], the reduction factor did not experience values
below 0.8. Despite different values of rc, it should be adjusted for the experimental data and type
of reinforcing.

10.9.3. Bond-slip models in ATENA

ATENA contains three models to simulate the bond and slip at the concrete-reinforcement
interface: the CEB-FIP model code 1990 [29], the bond-slip law proposed by Bigaj [30], and the
user defined bond model. The perfect bond (no slip) model can be defined in the user-defined
part also.

In the first two models (CEB-FIP and Bigaj models), the laws are created based on the
compressive strength of concrete, diameter and type of steel bar. The confinement conditions
and the quality of the concrete casting are other effective parameters in the definition of bond
behaviour in these models. In this study, of the three available bond models, the one best able to
simulate the observed behaviour of the test slabs was the user defined model. Therefore, the
experimental bond-slip model for LWC in Chapter 5 is used for this purpose. Also, the proposed
bond-slip models for LWC in Chapter 4 are used for comparison purposes. Figures (10.8) and
(10.9) show the table of input data to define the bond-slip model in ATENA 2D and 3D analysis
environments.

In the analysis of the slab in this study, the mechanical anchorage at both ends of the steel
bars, coupled with the deformed shape of the reinforcing bars proves to provide a good bond
between the cracked concrete and the longitudinal reinforcing bars.

401
Chapter 10- Numerical analysis

Figure 10.8: Reinforcement and bond details in ATENA (2014) 2D analysis [5]

10.9.4. Crack Models in ATENA

In the broad sense, the constitutive models of concrete may be classified into two categories:
the macroscopic phenomenological models and the micromechanical models. Two major
approaches are available in the literature for the modelling of cracking in concrete structures,
namely the discrete crack approach and the smeared crack approach.

The development of constitutive models that are capable of describing the behaviour of
concrete is a significant aspect in smeared crack modelling of concrete. ATENA includes two
smeared crack approaches to simulate the effects of cracking. The fixed crack and rotated crack
models are two types of smeared crack approach. Concrete without cracks is considered
isotropic, while concrete with cracks is orthotropic. The material axes of cracked concrete on
the orthotropic axes can be defined by two models; rotated or fixed cracks. Figure (10.10)
shows the fixed and rotated crack models in ATENA.

402
Chapter 10- Numerical analysis

Figure 10.9: Reinforcement and bond details in ATENA (2014) 3D analysis [5]

Fixed crack Rotated crack


Figure 10.10: Smeared crack models in ATENA [5]

Due to the assumption of isotropy in the concrete component, the directions of principal
strain and stress coincide in the uncracked concrete. After cracking the orthotropy is presented.
The weak axis of the material (m1) is normal to the crack direction and the strong axis of the
material (m2) is parallel with the cracks.

403
Chapter 10- Numerical analysis

In the rotated crack model, the crack direction of the principal stress always coincides with
the principal strain direction. In the fixed crack model, the crack direction and the material axes
are defined by the principal stress direction at the initiation of cracking.

A significant difference in the above approaches is in the shear model on the crack plane. In
the fixed crack model, a strain field rotation generates shear stress on the crack plane.
Consequently, the modelling of shear becomes necessary. In case of the rotated cracks, a shear
on the crack plane never appears and the shear model need not be employed [13].

In this study, both the fixed and rotated crack models are used and compared. The crack
direction is determined from the direction of principal stress at the instant of crack initiation.
This direction is then fixed and applied as the axis of the material orthotropy.

10.10. MODELLING OF REINFOCEMENT IN ATENA

Reinforcement can be modelled in two ways in ATENA. The first way is a discrete bar,
which is defined as a geometrical multi-linear object. It is embedded in solid objects. Its
geometry is defined independently of macro-elements. Thus, one reinforcing bar can be
embedded in any number of macro elements. The second way is a smeared reinforcement,
defined as a composite material, which has no effect on the geometrical model [5]. Gilbert and
Sakka (2010) [9] used discrete truss elements to model the embedded reinforcement in concrete
to study the strength and ductility of slabs.

Gribniak et al. (2010) [19] applied the reinforcement in the layers with the isotropic
quadrilateral FE with four integration points to model the reinforcement and concrete. They
studied the effect of mesh size on deformation of bridge girders.

10.11. CRITERIA TO SELECT PROPER ELEMENTS IN ATENA

The implemented elements in ATENA can be divided into three main groups:

- Plane elements for 2D, 3D and axisymmetric analysis;


- 3d solid elements;
- Special elements, which comprise elements for modelling external cable, springs, gaps,
etc.

Almost all existing elements in ATENA are constructed using the iso-parametric formulation
with linear and/ or quadratic interpolation functions. The interpolation function in a hierarchical

404
Chapter 10- Numerical analysis

manner is a considerable advantage of iso-parametric elements in ATENA. Despite the


application of this manner for plane quadrilateral elements in general, ATENA applies it for
plane triangular elements, 3D bricks, tetrahedral and wedge elements too. Together with the
interpolation function, the numerical integration scheme is another effective parameter in FEA.
It strongly affects the element properties by integrating the stiffness matrix and nodal forces.
ATENA applies the Gauss integration scheme to incorporate the majority of the elements.

10.11.1.Plane elements

10.11.1.1. Truss 2D and 3D Element


This group of elements is coded as CCIsoTruss <xx> for 2D analysis and CCIsoTruss <xxx>
for 3D analysis. Figure (10.11) shows the geometry of CCIsoTruss elements. This type of
element has constant strain along its length.

Figure 10.11: Geometry of CCIsoTruss elements [5]

10.11.1.2. Plane Quadrilateral Elements


This group of elements is coded as CCIsoQuad <xxxx> for 4 element nodes, to CCIsoQuad
<xxxxxxxxx> for 9 element nodes. Plane 2D, axisymmetric and 3D problems can apply this
type of element. Figure (10.12) shows the geometry of CCIsoQuad elements.

Figure 10.12: Geometry of CCIsoQuad elements (by 4 to 9 nodes) [5]

405
Chapter 10- Numerical analysis

10.11.1.3. Plane Triangular Elements


This group of elements is coded as CCIsoTriangle<xxx> for 3 element nodes, to
CCIsoTriangle<xxxxxx> for 6 element nodes. Plane 2D, axisymmetric and 3D problems can
apply this type of element. Figure (10.13) shows the geometry of CCIsoTriangle elements.

Figure 10.13: Geometry of CCIsoTriangle elements (by 3 to 6 nodes) [5]

10.11.2.3D Solid Elements

10.11.2.1. Tetrahedral elements

This group of elements is coded as CCIsoTetra<xxxx> for 4 node element, to


CCIsoTetra<xxxxxxxxxx> for 10 node element. Figure (10.14) shows the geometry of
CCIsoTetra elements.

Figure 10.14: Geometry of CCIsoTetra elements (by 4 to 10 nodes) [5]

10.11.2.2. Brick elements

This group of elements is coded as CCIsoBrick<xxxxxxxx…> for 8 up to 20 node element.


Figure (10.15) shows the geometry of CCIsoBrick elements.

406
Chapter 10- Numerical analysis

Figure 10.15: Geometry of CCIsoBrick elements (by 8 to 20 nodes) [5]

10.11.2.3. Wedge elements

This group of elements is coded as CCIsoWedge<xxxxxx…> for 6 up to 15 node element.


Figure (10.16) shows the geometry of CCIsoWedge elements.

Figure 10.16: Geometry of CCIsoWedge elements (by 8 to 20 nodes) [5]

10.11.3.Spring Element

This type of elements is used to model the spring-type boundary conditions such as the linear
relationship of the external forces on the boundary of structure with the associated displacement.
CCSpring, CCLineSpring and CCPlaneSpring are three figures of this element. All these
elements are derived from 2D or 3D formulation of the CCIsoTruss<xx> element.

CCSpring-2D and 3D elements model the spring-related condition at a point.

CCLineSpring- 2D elements model the spring-like boundary conditions along a line.

407
Chapter 10- Numerical analysis

CCPlaneSpring-3D element models the spring-like boundary conditions along a triangular


area.

Figure (10.17) shows the geometry of spring elements.

(a) (b)
Figure 10.17: Geometry of a) 2D CCSpring and CCIsoLineSpring and b) 3D CCSpring and CCPlaneSpring
[5]

Reinforcement can be modelled in two distinct forms: discrete and smeared. Discrete
reinforcement involves using a form of reinforcing bars and is modelled by truss elements. The
smeared reinforcement is a component of composite material and can be considered either as a
single (only one constituent) material in the element under consideration or as one of most such
constituents. The former case can be a special mesh element (layer); while, the latter can be an
element with concrete containing one or more reinforcements. In both cases the state of uniaxial
stress is assumed and the same formulation of stress-strain law is used in all types of
reinforcement.

Table (10.1) summarizes the available types and codes of elements in ATENA.

408
Chapter 10- Numerical analysis

Table 10.1: Available element types in ATENA [5]

Type Name Description


CCIsoBrick Isoperimetric brick element (hexahedron)
3d solid CCIsoWedge Isoparametric wedge element E.g.: CCIsoWedge<xxxxxx>
CCIsoTetra Isoparametric tetrahedral element E.g. : CCIsoTetra<xxxx>
CCIsoTriangle Isoparametric triangular element E.g.: CCIsoTriangle<xxx>
CCIsoQuad Isoparametric quadrilateral E.g.: CCIsoQuad<xxxx>
Four nodes quadrilateral element composed of two triangle isoparametric
CCQ10 elements. This element must be defined by at least four corner nodes.
E.g.: CCQ10<xxxx>
4 nodes quadrilateral element composed of two triangles. Four corner nodes
must define this element. The material model at this element is evaluated at
the element centre. The constitutive secant matrix evaluated at the element
CCQ10Sbeta
centre is used throughout the whole element to calculate internal forces of
element.
E.g.: CCQ10Sbeta<xxxx>
Plane Spring element defined by a single node. This element type should be used
CCSpring
to define a spring support at given node.
Line spring element defined by two nodes. This element type should be used
CCLineSpring
for spring supports along solid element edges.
Planar spring element defined by three nodes. This element type should be
CCPlaneSpring
used for spring supports along faces of solid elements.
Isoparametric truss element.
CCIsoTruss
E.g.: CCIsoTruss<xx>
Isoparametric truss element for axisymmetric problems. The element
contributes stiffness in the direction of its axis. Also, for adding radial
CCIsoASymTruss
stiffness, combine this element with the CCCircumferentialTruss or
CCCircumferentialTruss2 element. E.g. CCIsoASymTruss<xx>
CCIsoGap Gap/Interface element E.g.: CCIsoGap<xxxx>
Circumferential truss element. This element is defined by only one node and
is used in axisymmetric analysis to model circumferential reinforcement. It
CCCircumferentialTruss
also contributes radial stiffness.
E.g.: CCCircumferentialTruss
Circumferential truss element. This element is defined by two nodes and is
used in axisymmetric analysis to model circumferential reinforcement. It is
similar to the CCCircumferentialTruss element, however, it is “cross-
CCCircumferentialTruss2 sectional area” is equal to its length multiplied by its thickness. For adding
stiffness also in the element’s axial direction combine this element with the
Special CCIsoASymTruss element.
elements E.g.: CCCircumferentialTruss2
2D or 3D truss element to model the external prestress cables.
The bar is anchored at one end and prestressed at the other.
CCExternalCable The intermediate nodes are deviators, where frictional force is defined, see
external geometry definition. The whole bar must consist of one or more
elements. All the elements must compose the same element group.
2D or 3D truss element for modelling reinforcement bars with specified
cohesion with concrete. If exceeded, the bar will slip.
CCBarWithBond The element type uses external cable geometry definitions to specify the
appropriate solution parameters. The whole bar must consist of one or more
elements. All the elements must compose the same element group.

409
Chapter 10- Numerical analysis

Table 10.1: Available element types in ATENA [5] (Continued)

Type Name Description


3D shell elements. The first and the second digits in the element name
specify the number of integration points for element bending and shear
energy. The digit three notes that the element is integrated in 3 IPs in X-dir *
3 IPs in Y-dir * number of layers. The last letter L, H and S stands for 9-
CCAhmadElement33L nodes Lagrangian element, for 9 nodes Heterosis element and 8 nodes
CCAhmadElement32L Serendipity element. See theoretical manual for more details. All elements
CCAhmadElement33H must use a 3D material and LayeredShell geometry! They specified by 16
CCAhmadElement32H nodes, 8 for top and 8 for bottom surface similar to brick elements. The top
CCAhmadElement22S and bottom middle points for Lagrangian and Heterosis elements (for the
bubble functions) are generated automatically. At each node, the elements
have 3 degrees of freedom. As top and bottom node have altogether, 6 DOFs
and shell theory uses only 5 DOFs per shell node, the z-displacement of the
bottom node is automatically constrained during the execution.
3D non-linear beam element. The element uses quadratic interpolation along
its axis so that it can have a curvilinear shape. Similar to the implemented
CCBeamNL CCAhmad elements it is also input to a 3D hexahedral box. Nevertheless,
the usual axial nodal points are available (e.g. for checking resulting
deformations and rotations. They are generated automatically.

Table (10.2) explains the available material types in ATENA.

410
Chapter 10- Numerical analysis

Table 10.2: Material types in ATENA [5]

Material and sub-command Description


Linear elastic isotropic materials for 1D, Plane Stress, Plane Strain,
LINEAR_ELASTIC_ISOTROPIC
Axisymmetric and 3D analyses
3DCEMENTITIOUS Material suitable for rock or concrete like materials.
Materials suitable for rock or concrete-like materials. Enhanced
3DNONLINCEMENTITIOUS
3DCEMENTITIOUS material.
Materials suitable for rock or concrete like materials. This material is
3DNONLINCEMENTITIOUS2 identical to 3DNONLINCEMENTITIOUS except that this model is
fully incremental.
Materials suitable for rock or concrete-like materials. This material is
3DNONLINCEMENTITIOUS2VARI
identical to 3DNONLINCEMENTITIOUS2 except that selected
ABLE
material parameters can be defined using time or load step function.
Materials suitable for rock or concrete-like materials. This material is
3DNONLINCEMENTITIOUS2USER identical to 3DNONLINCEMENTITIOUS2 except that the selected
material laws can be defined by user curves.
Strain Hardening Cementitious Composite material. Material suitable
3DNONLINCEMENTITIOUS2SHCC
for fibre reinforced concrete, such as SHCC and HPFRCC materials.
Based on the 3DNONLINCEMENTITIOUS2 material, suitable for
3DNONLINCEMENTITIOUS2FATIGUE
fatigue analysis of rock or concrete like materials.
Materials suitable for rock or concrete-like materials. This material is
an advanced version of 3DNONLINCEMENTITIOUS2 material that
3DNONLINCEMENTITIOUS3 can handle the increased deformation capacity of concrete under
triaxial compression. Suitable for problems including confinement
effects.
VON_MISES_PLASTICITY Plastic materials with Von-Mises yield condition, e.g. suitable for steel.
DRUCKER_PRAGER_PLASTICITY Plastic materials with Drucker-Prager yield condition.
User-defined material (derived from elastic isotropic). The user
USER_MATERIAL
provides a dynamic link library.
INTERFACE_MATERIAL Interface material for 2D and 3D analysis.
REINFORCEMENT Material for discrete reinforcement.
REINFORCEMENT_WITH_
Material for discrete reinforcement subject to cycling loading.
CYCLING_BEHAVIOR
SMEARED_REINFORCEMENT Material for smeared reinforcement.
MICROPLANE Bazant Micro-plane material models for concrete
This material can be used to create a composite material consisting of
various components, such as for instance concrete with smeared
reinforcement in various directions. The unlimited number of
COMBINED_MATERIAL
components can be specified. Output data for each component are then
indicated by the label #i; where i indicates a value of the i-th
component.
BEAM_RC_MATERIAL Material for (reinforced) structures modelled by CCBeal material
Material for reinforcement bar used in solids modelled by either
try_reduce_MyMz_keep_NBEAM_R
BEAM_RC_MATERIAL or BEAM_MASONRY_MATERIAL
EINF_BAR_MATERIAL
material.

10.12. CREEP AND SHRINKAGE ANALYSIS

The time-dependent analysis of the reinforced concrete structures implements creep and
shrinkage data. The main difference between the usual static and creep analysis is the inclusion
of the structural response integration in time, in the creep analysis. Time factor also appears in

411
Chapter 10- Numerical analysis

the constitutive equations such as material models by implementing the models to predict the
creep and shrinkage behaviour of concrete.

There are different models of creep used in ATENA to analyses the time-dependent cracking
and deflection of reinforced concrete member. Table (10.3) shows the existing creep models in
ATENA used to assign to the concrete base material in finite element analysis.

Table 9.3: Material models for creep in TANEA [5]

Material Description
CCModelB3 Bazant-Baweja B3 model
CCB3Improved Model same as the above with support for specified time and humidity history
CCModelBP_KX Creep model developed by Bazant-Kim, 1991
CCModelCEB-FIP Creep model advocated by CEB-FIP 1978
CCModelACI_78 Creep model by ACI Committee in 1978
CCModelCSN731202 Model recommended by CSN731202
CCModelBP1 Full version of the creep model developed by Bazant-Panulla
CCModelBP2 Simplified version of the above model
Creep model for direct input of material compliance, strength and shrinkage at
CCModelGeneral
times typically measured in a laboratory
CCModelFIB_MC2010 Model by CEB-fib bulletin 65 from the year 2010
CCModelEN1992 Creep model by Eurocode EN1992.1.1-2006

Completeness and continuity are two aspects of the so-called consistency condition between
the discrete and mathematical models, while, consistency and stability imply convergence in
FEA.

In FEA, it is crucial to select an element that is compatible with the assigned element type.
Element types that are incompatible with the material type of the concrete member may cause
some convergence problems in solving the non-linear equations. For a structural FE, the
stiffness matrix includes the material behaviour and geometric information that points to the
resistance of the element to deformation when subjected to loading. The shape functions of the
element should provide displacement continuity between elements to ensure that no material
gaps appear as the elements deform.

Table (10.4) explains the compatibility between the type of elements and the chosen materials
in FEA in ATENA.

412
Chapter 10- Numerical analysis

Table 10.4: Element type and material compatibility [5]

CCCircumferent
CCPlane-Spring
CCIso-Triangle

CCLine-Spring
CCQ10Sbeta

CCIso-Truss
CCIso-Quad
CCIso-Brick

CCIso-Gap
Material

CCSpring

ialTruss
CCQ10
CC1DElastIsotropic ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬

CCPlaneStressElastIsotropic ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬

CCPlaneStrainElastIsotropic ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬

CC3DElastIsotropic ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬

CCASymElastIsotropic ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬

CC3DBiLinearSteelVonMises ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬

CC3DCementitious ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬

CC3DNonLinCementitious ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬

CC3DNonLinCementitious2 ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬

CC3DNonLinCementitious2User ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬

CC3DNonLinCementitious2Variable ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬

CCSBETAMaterial ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬

CC3DDruckerPragerPlasticity ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬

CC2DInterface ‫ݱ‬

CC3DInterface ‫ݱ‬

CCReinforcement ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬

CCCyclingReinforcement ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬

CCSmearedReinf ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬

CCCircumferentialSmearedReinf ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬

CCMaterialWithVariableProperties ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬

CCMaterialWithTempDepProperties ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬

CCMaterialWithRandomFields ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬

CCCombinedMaterial ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬

10.13. MODELIING OF SHRINKAGE BEHAVIOUR IN ATENA

There are three different methods in TAENA to implement the shrinkage effect in FE models:

- Reducing the tensile strength of concrete and/ or reducing the fracture energy. This
option takes the reduction of tensile strength due to cracking of concrete;

413
Chapter 10- Numerical analysis

- Introducing the shrinkage as initial strain due to constant volume reduction throughout
the whole structure volume. This type of initial constant strain is typically in the range of
150 to 450 micro-strains;
- Applying the real shrinkage data of concrete in creep analysis module. Since this option
takes into account the exact shrinkage data, it gives the most accurate FEA results;

Figure (10.18) shows a typical shrinkage data input box in ATENA-CREEP analyses.

Figure 10.18: Shrinkage input data implementation in ATENA-CREEP analysis

10.14. SELECTION OF COMPATIBLE ELEMENTS WITH


EXPERIMENTAL DATA

The obtained results from experimental studies have been verified by 2-D and 3-D FE
analysis in the GID-ATENA program.

The monitored parameters of analysis of one-way slabs in 2D and 3D environments were


deflection and cracking behaviour under instantaneous loading conditions. Additionally, the
time-dependent effects of creep and shrinkage on the long-term deflection over 500 days were
investigated. The cracking pattern along the slab span and the time-dependent deflection
variation were also investigated in both 2D and 3D models. According to the experimental
program, two different levels of instantaneous and sustained loading were applied on the slabs
in 2D and 3D models to monitor the cracking and deflection behaviour when subjected to
instantaneous and time-dependent behaviour.

414
Chapter 10- Numerical analysis

Based on the previously explained properties of materials and element types implement in
ATENA, the quadratic element type in Table (10.4) was selected in the 2-D analysis. Compared
to the analysis of other available element types, the selected element showed the best
compatibility with the experimental data of instantaneous deflection and behaviour.

For 3D analysis, CCIsoBrick element in Figure (10.15) was the most appropriate element to
cover the analysis objectives and applied material types in this study. To decrease the
computation time of the model processing, the CCIsoBrick element was configured in 8 nodes.
Considering the higher accuracy of the 20-nodded CCIsoBrick element compared to 8-nodded
element in the model, to compensate for the reduced efficiency of 8-nodded element, a finer
structural mesh size together with the quadratic type option have been utilized in meshing the
model before processing. The finer structural mesh size and 8-nodded CCIsoBrick element
provide an optimized analysis in terms of the processing time and required efficiency.

The effect of mesh size on the FE analysis results is described in the next section of this
study.

10.15. EFFECT OF MAXIMUM AGGREGATE SIZE AND AGGREGATE


INTERLOCK

The aggregate interlock transfers the shear loads across the crack faces. It is reliant upon the
frictional property of the aggregates along the crack faces and the crack opening. Type and
maximum size of the aggregate, crack tortuosity, stiffness and strength of concrete, boundary
conditions and loading characteristics affect the transferred shear throughout the crack surface
due to the aggregate interlock [31, 32].

According to Figure (10.19), in GID-ATENA, it is possible to include the effects of


aggregate interlock and maximum size of coarse aggregate on the shear and compressive
strength of the reinforced concrete structure in FE analysis. For realistic analysis and the most
accurate simulation of the experimental conditions, the aggregate interlock is activated during
the analyses. Also, the maximum size of aggregate is set to 10 mm. The aggregate size is in
agreement with the experimental phase of this study.

415
Chapter 10- Numerical analysis

Figure 10.19: Implementation of aggregate interlock and maximum size of aggregate in FEA in ATENA

10.16. EFFECT OF MESH SIZE

The mesh quality of FE has a very significant influence on the quality of the analysis results.
The size and type of the mesh affect the FEA in the following ways:

- Considering the accuracy of the FEA, proper selection of size and type of mesh is an
effective way to make substantial savings in computational time and computer memory;
- The pattern of crack propagations depends on the main directions of the mesh. However
the problem is less important in discrete elements [10];
- Strain localization in the post-peak softening of concrete is greatly influenced by the
mesh size and type;
- In a FEA, the overall non-linear response of a reinforced concrete member subjected to
bending can significantly be affected by the mesh size and type. This effect is more
critical in using the smeared crack modelling [33];
- Compatibility of the mesh size and type in contact surfaces like the simulation of the
steel-concrete bond is inevitable in convergence and analysis of the model.

The effect of FE mesh sizes on the analysis of the instantaneous and time-dependent
behaviour of slabs under two different load levels is investigated. Displacements at midspan and
third point and computational time are the two main categories in this regard.

416
Chapter 10- Numerical analysis

10.17. WIDTH AND HEIGHT OF THE SELECTED ELEMENTS

Utilization of each type of element in GID-ATENA requires the specified quantity or width
and height of element in each direction of the model to get the most efficient models of
structures by FE analysis.

In FE analysis, the faces of 2-D elements build the 3-D elements. The CCIsoBrick element
was utilized in the 3-D analysis of the slabs and CCIsoQuad in Table (10.4) is its plane version.
Considering the thickness of slabs and the position of the reinforcement layer, the minimum
number of 8 elements through the thickness of the slab is selected. The element type was
CCIsoQuad. In other words, the minimum dimension of CCIsoQuad element throughout the
slab thickness (161 mm) was selected as 20mm. The quantity of 6 to 8 finite elements per height
demonstrates sufficient accuracy models of reinforced concrete slabs.

10.18. EFFECT OF LOADING PADS IN PROGRAMMING

Loading pads are interface elements between the heavy loading blocks and the main structure
to be loaded. The loading pads ensure that the load transfer is correct through the width of the
pad. Therefore, the loading pads eliminate any chance of fretting at the edge due to the rotation
of the slab.

The interface between the loading pads and the slab is a type of contact element. Generally,
in the iterative scheme, the contact conditions in FEA model are classified into three groups:
[34]

- Open condition; gap remains open;


- Stick condition; gap remains closed, and no sliding motion occurs in the tangential
direction;
- Sliding contact; gap remains closed, and the sliding takes place in the tangential
directions.

The loading pads transfer the applied loading to the slab, however the stiffness of the pad is
transmitted into the slab as displacement of the mesh within each element is linked to the
contact surface. This stiffness transition causes the localized stress and strain concentration, due
to the increased stiffness in some areas.

To simulate the real behaviour of the experimental program, the pads under the loading block
are also implemented in the slab models. Figure (10.20) shows the loading pads in real test

417
Chapter 10- Numerical analysis

specimens and the simulated pads in GID-ATENA program. These pads are modelled in both
2D and 3D models.

Figure 10.20: Simulated and real loading pads

10.19. ADVANTAGES OF SLABS SYMMETRY IN NUMERICAL


ANALYSIS

The investigated slabs are in the shape of a long prism, and the position and condition of
supports, reinforcement arrangement and loading pads distribution are symmetric all over the
slabs. Therefore, it is possible to analyse half of the model to save the processing time, keeping
the mesh quality high and achievable efficiency of the model.

To simulate the whole slab behaviour by a half slab, the boundary conditions in the FEA
should be set properly. According to the slab geometry and support conditions, there is no
rotation at the mid-span cut section. Also, the lateral transition in “z” direction and longitudinal
movement along the span length in “x” direction in that section are supposed to be zero.

The tensile stress at the extreme point of the half-slab (mid-span of whole slab) is a crucial
monitoring point in this study. Hence, the CCIsoBrick element in Table (10.2) can be used with
high accuracy to mesh the half-slab model. The optimum width and height of the element are
defined based on the sensitivity of FEA to the mesh size in 3D analysis.

Figure (10.21) shows the general view of the half-slab model in GID environment. This
model is the same for all four slabs in this study.

Figure 10.21: Half-slab model of slab in GID environment


418
Chapter 10- Numerical analysis

GID-ATENA is a FE-based software system explicitly developed for the non-linear analysis
of reinforced concrete structures. ATENA is used for the analysis of the model, while the
program GiD is used for the data preparation, modelling of the structure and the mesh
generation.

GID is a universal, adaptive and user-friendly pre and post-processor for numerical
simulations in science and engineering. It has been designed to cover all the common needs in
the numerical simulations field from pre to post-processing including: geometrical modelling,
effective definition of analysis data, meshing, and data transfer to analysis software, as well as
the visualization of numerical results. The main advantage of using GID, together with ATENA
is the reduced analysis time with higher quality of meshing for the simulated reinforced concrete
structure.

In the pre-processing stage, GID is used to define the geometry of the structure, material
properties, boundary conditions and analysis data including the loading steps, solving method of
non-linear equation and the convergence method.

Three-dimensional regions are modelled by volumes in GiD. The embedded reinforcement is


modelled as a line and is not usually connected to any surface or volume.

10.20. ANALYSIS OF SLABS UNDER INSTANTANEOUS LOADING

10.20.1.Process of the modelling

In the two-dimensional environment of GID-ATENA, two types of analysis have been


applied on the simulated models of LWC slabs. The first type is non-linear static analysis, to
study the cracking behaviour and deformation of slabs under the instantaneous load on day zero.
The day zero, is the first day that the slabs were loaded. In other words, this type of analysis
doesn’t take into the account the time-dependent properties of concrete.

The study of cracking behaviour study in this research includes crack propagation and
direction under instantaneous loading. Deflection study includes the maximum flexural
deflection at the mid-span and, the deflection in the third point under instantaneous loading.

The version of the GID program used to model the structure in this study is 12.0.6 and the
version of ATENA program linked to this GID program to analysis the models is V5.

419
Chapter 10- Numerical analysis

10.20.2.Slab dimensions

Mesh size and number of elements are two crucial parameters in the FEA of reinforced
concrete structures. Finer mesh size and higher number of elements provide a more accurate
analysis of structure. Considering the proper mesh size, whole structure with real boundary
conditions, loading pads and supporting pads is modelled in FE and analysed subjected to two
different loading levels. Figure (10.22) shows the details and geometry of the slab model built in
the GID program and the applied mesh size. Two points in the mid-span and the third point are
monitored for instantaneous deflection.

Figure 10.22: Details, geometry and meshing of the LWC slabs in GID program

10.20.3.Basic material properties

According to the experimental results presented in Chapter 5, Tables (10.5) and (10.6)
explain the basic material properties of concrete and reinforcing steel bars implemented in the
FE models.

Table 10.5: Mechanical properties of lightweight concrete

Age
Property
14 days 28 days

Strength class in ATENA 25/30 30/37

Elastic modulus (GPA) 17.25 19.48

Compressive strength (MPa) 27.93 30.89

Tensile strength (MPa) 2.6 2.89

Strain at peak stress 0.000098 0.00011

Type of tension stiffening Exponential Exponential

Crack model Fixed Fixed

420
Chapter 10- Numerical analysis

Table 10.6: Mechanical properties of steel bars

Property Value or type

Elastic modulus (GPA) 200.553

Yield strength (MPa) 538

Hardening Perfectly plastic


3
Density (kg/m ) 7850

Additionally, the proposed models of material properties in Chapter 6 are utilized in ATENA
models and the results are compared.

According to the existing material models in ATENA in Tables (10.1) and (10.2), the
following models can be used to simulate the concrete properties in this study.

- CC3DCementitious material:

This material model assumes a linear response of compressive and tensile concrete, up to the
point when the failure envelope is reached. This material model doesn’t consider the hardening
regime.

- CC3DNonLinCementitious material:

In contrast with CC3DCementitious, this material model assumes a hardening regime before
the compressive strength is reached.

- CC3DNonLinCementitious2 material:

Except the purely incremental formulation, this material model is equivalent to


CC3DNonLinCementitious. Therefore, it is proper to apply for the creep calculation, or when
the properties of material require some modification during the analysis.

- CC3DNonLinCementitious2Variable material:

In addition to the included properties in CC3DNonLinCementitious2, this model contains the


history evaluation law’s definition of the material parameters.

- CC3DNonLinCementitious2User material:

In this model, the user can define the compression, tensile, shear and softening diagrams and
all other required properties of the concrete.

Considering the existing models in GID-ATENA, CC3DNonLinCementitious2 model of


concrete is used for a general analysis. It allows changing the material properties during analysis
421
Chapter 10- Numerical analysis

and can consider the creep effect also. Figure (10.23) shows the table of implemented properties
of this material model in ATENA. Some implemented values of the tables in the figure are the
default values of the GID-ATENA program.

To model the reinforcement material in the FE model, a multi-linear law consisting of four
lines is utilized to adopt to the experimental stress-strain test result of steel bars in Chapter 6.
The implemented multi-linear law contains the elastic state, yield plateau, hardening state and
the fracture point.

Considering two types of smeared and discrete reinforcement models in GID-ATENA, the
stress-strain law is used to describe both the smeared and discrete models. However, to model
the smeared reinforcement model, the reinforcement ratio in the section and angle of direction is
required to implement in ATENA input data.

Figure 10.23: Material properties for CC3DNonLinCementitious2

422
Chapter 10- Numerical analysis

Figure (10.24) shows the typical input data table to define the reinforcement in GID-ATENA.
The table includes the stress-strain curve definition, and the bond-slip behaviour between the
reinforcement and surrounded concrete.

Figure 10.24: Parameters to define the reinforcement in GID-ATENA

Despite constant values of the basic parameters of concrete and reinforcement, there is wide
range of parameters to define their other mechanical properties in GID-ATENA.

In the concrete material definition part of the ATENA program, the variable parameters in
tensile, compressive and shear section are presented in Table (10.7). To find out about the most
compatible adjustment of the parameters with the experimental results, different arrangements
of the effective parameters in tensile, compressive and shear partition in the input data have
been examined. In addition, the sensitivity of the model to various values of other constant
parameters in tension, compression and shear has been investigated.

Table 10.7: Variable parameters in definition of typical concrete properties in GID-ATENA [5]

Property Type of Tension Softening (TTS) Crack model


Local SFRC local
Tensile Exponential Linear SFRC Fixed Rotated
strain strain

Type of Compression Softening (TCS)


compressive Crush band Softening modulus

Shear Retention Factor (SRF) Tension-compression interaction (TCI)


Shear Fixed Variable Linear Hyperbola A Hyperbola B

423
Chapter 10- Numerical analysis

10.20.4.Solution parameters in instantaneous loading analysis

Table (10.8) describes the applied solution method and parameters of converging the analysis
of the models subjected to instantaneous loading. The applied parameters provide the required
convergence and the most accurate results with a reasonable processing time of the models.

Table 10.8: Solution parameters in instantaneous deflection analysis of LWC slabs

Solution method Newton-Raphson

Stiffness/update Tangent/each iteration

Number of irritations 80

Error tolerance 0.01

Line search On, with iteration

Optimization Sloan

Element shape smoothing On

Figure (10.25) shows a typical convergence diagram of the slab analysis in GID-ATENA
program.

Figure 10.25: Typical convergence of the model analysis in ATENA program

424
Chapter 10- Numerical analysis

According to Figure (10.25), there are four convergence criteria to solve the non-linear
equations of the model.

Criteria 1: relative error in displacement

Criteria 2: relative error in residual forces

Criteria 3: absolute error in residual forces

Criteria 4: relative error in energy, i.e. basically a product of relative residual error and
displacement error

10.20.5.Monitoring points of instantaneous deflection

The maximum deflection at mid-span and the deflection at the third point of the slab length
were the main investigated parameters in the instantaneous loading of the slabs. Furthermore,
the cracking behaviour of the slabs under two different levels of uniformly distributed loading
has been studied. The ratio of the deflections at monitoring points is a matter of investigation as
well.

There are two different levels of uniformly distrusted loading on the simply supported slabs.
The loading values of 30% and 40% of the ultimate loading capacity of the slab section were
applied to each pair of slabs. In the experimental program, the loading levels of 0.3Mu and
0.4Mu are described as loading level 1 (LL1) and loading level 2 (LL2) respectively in all model
analyses.

According to the experimental data in Chapter 6, the recorded instantaneous deflection at the
monitoring points in two loading levels is presented in Table (10.9).

Table 10.9: Experimental instantaneous deflection values under LL1 and LL2 in monitoring points

Loading level (1) Loading level (2)


Slab Δ1(mm) Δ2 (mm) Δ1(mm) Δ2 (mm)
LWC-1 11.84 11.19
LWC-2 10.76 9.22
LWC-3 6.33 5.43
LWC-4 5.98 5.16

As part of calibration of the ATENA program to analyse the LWC slabs, Table (10.10)
describes different arrangements of the variables in the definition of SBETA material type.

425
Chapter 10- Numerical analysis

Also, the obtained deflections in the monitoring points of the slab under loading level (1) are
presented in the table.

Table 10.10: Variable arrangement of parameters in SBETA material and the obtained deflections in LL1

Tensile Compressive Shear Deflection (mm)


Case TTS Crack model ࢌᇱࢉ reduction due to cracks TCS SRF TCI Point 1 Point 2
1 SFRC Fixed 0.8 0.2 Fixed Linear 5.98 5.2
2 exponential Fixed 0.8 0.2 Fixed Linear 11.8 10.3
3 linear Fixed 0.8 0.2 Fixed Linear * *
4 local strain Fixed 0.8 0.2 Fixed Linear 11.9 10.4
5 SFRC local strain Fixed 0.8 0.2 Fixed Linear 11.9 10.4
6 SFRC Rotated 0.8 0.2 Fixed Linear 5.98 5.19
7 exponential Rotated 0.8 0.2 Fixed Linear * *
8 linear Rotated 0.8 0.2 Fixed Linear 11.5 10
9 local strain Rotated 0.8 0.2 Fixed Linear 11.8 10.3
10 SFRC local strain Rotated 0.8 0.2 Fixed Linear 11.8 10.3
11 SFRC local strain Rotated 0.7 0.2 Fixed Linear 11.8 10.3
12 SFRC local strain Rotated 0.6 0.2 Fixed Linear 11.8 10.3
13 SFRC local strain Rotated 0.5 0.2 Fixed Linear * *
14 SFRC local strain Rotated 0.4 0.2 Fixed Linear 11.8 10.3
15 SFRC local strain Rotated 0.2 0.2 Fixed Linear 11.8 10.3
16 SFRC local strain Rotated 0 0.2 Fixed Linear * *
17 SFRC local strain Rotated 0.3 0.2 Fixed Linear * *
18 SFRC local strain Rotated 0.1 0.2 Fixed Linear 11.8 10.3
19 SFRC local strain Rotated 0.8 0.2 Fixed Linear 11.8 10.3
20 SFRC local strain Rotated 0.8 0.2 Fixed Linear 11.8 10.3
*
Model processing failed

After calibrating the model with acceptable instantaneous deflection in agreement with the
experimental results, other parametric studies have been performed, and the results are
presented.

10.20.6.Effect of mesh size on deflection and cracking of slabs

After selecting the most compatible element type to allow for study of the effect of mesh size
on the cracking pattern, the maximum size of cracks and the deflection of slabs, six distinct
sizes of elements by meshing have been implemented in the finite element model. Except the
changing mesh size, all other parameters such as material properties, loading values and support
conditions were kept constant during the analysis. Table (10.11) explains the applied type of
element and other preferences in the model analysis with different mesh sizes.

Table 10.11: Preferences in studying the effect of mesh size on instantaneous deflection

Parameter Mesh type Element type Loading steps Iteration number Bond-slip model

Preference Quadrilaterals CCIsoQuad 5 40 Perfect connection

426
Chapter 10- Numerical analysis

Figure (10.26) shows the deflection and crack propagation at the critical flexural area of
mid-span by processing the model with different mesh sizes in 2D environment.

7 mm mesh
Figure 10.26: Crack propagation and deflection in the mid third area of the slab span by different mesh sizes

Figure (10.27) shows the crack propagation along the span length by applying different mesh
sizes.

Figure 10.27: Cracking pattern and propagation in slab spans by different mesh sizes

According to Figures (10.26) and (10.27), the mesh size changes the propagation and
direction of the cracking in the high moment region (middle third of span length) critical
flexural range of the slab and whole slab length. It also changes the deflection curve of the slab
under uniformly distributed loading and the maximum deflection of the slab at mid-span.

10.20.7.Effect of bond-slip model

To study the effect of bond-slip effect on the instantaneous deflection of slabs, the perfect
bond between the steel bars and concrete is compared with the existing bond-slip models in
427
Chapter 10- Numerical analysis

ATENA. Figure (10.28) describes the recorded maximum deflection of models with different
mesh size to the maximum deflection of the model by 7 mm mesh. There were no considerable
changes in the values of deflection and crack width in the models by finer mesh size, therefore,
the 7 mm was selected as the basis for evaluation of the effect of bond-slip model. According to
Figure (10.28) the impact of different bond-slip relationships on the models by different mesh
sizes is evident.

104 P1 (perfect bond)


P2 (perfect bond
P1 (Malvar et al, 2003)
100
P2 (Malvar et al, 2003)
P1 CEB-FIP,1993)
96 P2 (CEB-FIP,1993)

92
Deflection ratio (%)

88

84

80
80 50 20 10 8 7
Mesh size (mm)
Figure 10.28: Ratio of deflection by different mesh sizes to deflection by 7 mm mesh

P1 and P2 in Figure (10.28) are deflections of slab at mid-span and the third point of span
length respectively.

To study the effect of bond-slip behaviour on the maximum crack width under the
instantaneous loading, the model with perfect bond is compared with the models with
implemented bond-slip models in ATENA. Figure (10.29) shows the ratio of the maximum
crack width in models with different mesh sizes to the maximum crack width in a model with 7
mm mesh size. There is no considerable difference in the maximum crack width of the models
at midspan by perfect bond or bond-slip relationship between the steel bars and concrete.

Similar results were obtained for the monitoring point (P2) at the border of third middle of
slab length; i.e. the inclusion of perfect bond or bond-slip between the steel bars and concrete
did not show considerable changes in the crack width at the third point of slab length under
428
Chapter 10- Numerical analysis

instantaneous loading. However, the maximum crack width in mid-span was about 6% smaller
than the maximum crack width in the third point by analysis of models with different mesh
sizes.

10
Perfect bond

8 CEB-FIP 1993
Ratio of crack width Malvar et al, 2003
6

0
80 50 20 10 8 7
Mesh size (mm)

Figure 10.29: Ratio of the maximum crack with at different mesh sizes to crack with by 7 mm mesh

10.20.8.Ratio of deflection at monitoring points

Despite different values of deflection at mid-span (Δ1) and deflection at the third point of slab
length (Δ2) in the models with perfect bond and bond-slip model, the ratio of Δ1/Δ2 is
approximately constant.

The maximum crack widths at the monitoring points experience a similar behaviour. Despite
different values of the maximum crack width at mid-span (Wmax-1) and the maximum crack
width at the third point of slab length (Wmax-12) in the models with perfect bond and bond-slip
model, the ratio of wmax-1/ wmax-2 is approximately constant.

Inclusion of the bond-slip behaviour between the steel bars and concrete in the model
changes the maximum deflection value at the mid-span and the maximum crack width.
However, the changes are not considerable. In terms of deflection, in the model with perfect
bond (no slip) between the steel bars and concrete, the maximum deflection at mid-span is about
4% less than the corresponding deflection in a model with bond-slip behaviour. However, this
ratio may change by applying different bond-slip relationships in the model.

429
Chapter 10- Numerical analysis

10.20.9.Effect of support condition on the instantaneous deflection

Different support conditions are investigated to monitor the ratio of Δ1/Δ2. The ratio of
deflection in the corresponding monitoring points of Simply Supported (SS) slab to the Fixed
Ends (FE) slab is given in Table (10.12).

Table 10.12: Ratio of deflection in corresponding points of simply-supported and fixed ends slab

Parameter Δ1-SS/ Δ2-SS Δ1-FE/ Δ2-FE Δ1-SS/ Δ1-FE Δ2-SS/ Δ2-FE

Non-linear 1.17 1.22 2.88 3.02

According to Table (10.12), fixing the slab ends reduces the maximum instantaneous
displacement of mid-span to 0.33 of the maximum deflection in a simply supported slab in the
range of service load. However, the maximum crack width in the negative moment (fixed
support) zone, especially in the slab with one-layered steel bars in tension zone, is considerably
higher, compared to the maximum crack width in the simply-supported beam.

10.20.10. Sensitivity of deflection to early age properties of concrete

Conventionally, the mechanical properties at age 28 days are considered as the


representatives of the concrete material. The slabs in the experimental part of this study have
been subjected to loading at age 14 days. To investigate the effect of concrete properties at
different ages on the instantaneous cracking and deflection behaviour, the slab model is
analysed by implementing the concrete properties at 14 and 28 days.

The measured values of compressive strength and modulus of elasticity at the age of 14 days
are 90% and 87% of those values at age 28 days respectively. However, in the range of service
load, deflection of the monitoring points at 14 days is 10% higher than those at 28 days.
Surprisingly, the non-linear static analysis of the slab showed about 30% higher value of the
maximum crack width at 14 days under the same support conditions and applied load values.

Deflection (Δi-t) and the maximum crack width (wmax-t) ratios of the simply supported LWC
slab in the monitoring points, subjected to instantaneous loading by different properties of
concrete at 14 and 28 days, is presented in Table (10.13).

430
Chapter 10- Numerical analysis

Table 10.13: deflection ratio at the corresponding point by concrete properties at 14 and 28 days

Δ1-14*/ Δ1-28 Δ2-14/ Δ2-28 Wmax-14/Wmax-28

Deflection ratio 1.1 1.1 0.73

In Table (10.13), Δi-d is deflection at monitoring point “i” with properties of concrete at the
age of “t” days. Wmax-d is the maximum crack width in the model with properties of concrete at
age “t” days.

10.20.11. Effect of loading pads

In the FE models in this study, the loading pads have been implemented to simulate the real
experimental conditions. According to the experimental conditions, the applied concrete blocks
on the loading pads are heavy and huge, compared to the dimension of pads. Furthermore, the
contact surface between the pads and the rough top surface of the slab was filled and prepared
by plaster. Therefore, the contact condition between the loading pads and the slab surface is
assumed to behave together; i.e. the loading pads are perfectly bonded to the slab.

Figure (10.30) shows the non-uniform distribution of the stress and strain around the contact
area in different loading steps of the static non-linear analysis. The pads are modelled as flexible
plates. There is no crack initiation and crack propagation in the first load step. However, the
displacement value and the maximum crack width are increasing at higher loading values in
higher load steps.

Load step εyy σyy Von Mises stress

First

Last

Figure 10.30: Localized stress and strain distribution around the loading plate (flexible plates)

431
Chapter 10- Numerical analysis

Since the pads under the loading blocks were made of timber, the corresponding modulus of
elasticity and mechanical properties of timber were assigned to the pads in the analysis of the
slabs. However, some studies consider these types of plates as rigid plates in contact with the
slab. Therefore, to monitor the effect of the elastic or rigid pad on the instantaneous deflection
of slabs, different values of rigidity were assigned to the pads during the analysis of the model.

Figure (10.31) shows the corresponding parameters (localized stress and strain distribution)
presented in Figure (10.30) for a rigid loading pads.

Comparison of the flexible and rigid loading pads in Figures (10.30) and (10.31) shows a
considerable difference in the cracking pattern and distribution of the stress and strain around
the pads in the tensile zone of the concrete. In case of the rigid pad, there are more and wider
cracks around the pads due to higher levels of strain and stress transmitted from the rigid plate
to the slab.

Load step εyy σyy Von Mises stress

First

Last

Figure 9.31: Localized stress and strain distribution around the loading plate (Rigid plates)

432
Chapter 10- Numerical analysis

10.21. TIME-DEPENDENT ANALYSIS OF SLABS

The second type of analysis in GID-ATENA is the creep analysis to include the effects of
creep and shrinkage on the time-dependent cracking and deflection behaviour of slabs. Since the
experimental program has lasted for 500 days, the creep test was performed to monitor the slab
behaviour from the first day of loading up to 500 days monitoring the effects of creep and
shrinkage under the same sustained loading.

10.21.1.Modelling of the slab

Modelling of the structure for any type of problem is performed in the graphical environment
of GID. However, there are different features and environments of ATENA to analyses the
different models for various problems types. Figure (10.32) shows the pre-processing
environment in GID and the implemented problem types in ATENA to analyse the model. In
switching between the methods (problem types), it is possible to convert the data information
(material properties, support and loading conditions and other analysis data) of the existing
model to the new problem type.

According to Figure (10.32), there are different post-processing environments to analyses the
model. To perform the creep analysis, the modelling process is similar to the process of static
analysis. To process the model for time-dependent analysis, the problem type should be changed
to ATENA-Creep.

Figure 10.32: Different problem types in ATENA to analyse the model

Since the modelling of slab structures for both ATENA-Static (instantaneous loading
analysis) and ATENA-Creep (time-dependent analysis) problem types are accomplished in

433
Chapter 10- Numerical analysis

GID, the major part of the modelling process and preferences in 2D and 3D environment are
similar. The main differences between the pre-processing stage in static and creep analysis in
GID-ATENA are the implementation of the creep behaviour in the material properties and the
addition of the interval data (loading step) for creep application.

10.21.2.Basic material properties

Based on the calibration process of selecting the most compatible material in the model,
CC3DNonLinCementitious2 material model (Cementitious2) is chosen as the base material for
the solid concrete in the creep analysis of the slabs under different loading conditions.

The creep behaviour of concrete in the model can be implemented in direct or indirect ways.
In the direct method, the user can develop the creep model of concrete and specify all other
related details manually. However, there are some previously developed creep models
implemented in the ATENA program to facilitate the creed application in the model. Figure
(10.33) shows the existing creep models in ATENA.

Figure 10.33: Implemented creep models in ATENA-creep problem type

The creep models in Figure (10.33) have been developed for different conditions of physical
and mechanical properties of concrete, mixture proportioning, environment conditions and
loading history. Therefore, they may vary in accurately predicting the creep behaviour and the
required parameters of the model.

Table (10.14) describes the required parameters for each type of the creep models in ATENA.

434
Chapter 10- Numerical analysis

Table 10.14: Input parameters in the creep and shrinkage models implemented in ATENA

B3 improved

General
BP-KX

CEB

CSN
ACI

BP2
BPI
Model name B3

Concrete type ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬


Thickness (Volume/Surface) ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬
Strength fcyl28 ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬
Strength fcyl028 ‫ݱ‬
Fracture energy Gf28 ‫ݱ‬
Tensile strength ft28 ‫ݱ‬
Young modulus Ec28 ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬
Ambient humidity (h) ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬
Aggregate / cement ratio ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬
Water / cement ratio ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬
Fine/total aggregate ratio (sa) ‫ݱ‬ ‫ݱ‬
Coarse/fine aggregate ratio (gs) ‫ݱ‬ ‫ݱ‬
Fine aggregate/cement ratio (sc) ‫ݱ‬ ‫ݱ‬
Shape factor ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬
slump ‫ݱ‬
Air content ‫ݱ‬
Cement mass ‫ݱ‬
Concrete density ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬
Curing type ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬
End of curing ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬
Thermal expansion coefficient (αt) ‫ݱ‬
Autogeneous shrinkage at infinity time (εa,∞) ‫ݱ‬
Half-time of autogeneous shrinkage (τa) ‫ݱ‬
Time of final set of cement (ts) ‫ݱ‬
Final self-desiccation relative humidity (ha,∞) ‫ݱ‬
Current time t ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬
I/D
Load time ‫ ݐ‬ᇱ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬
Total water loss w ‫ݱ‬ ‫ݱ‬ ‫ݱ‬
Water loss w(t) ‫ݱ‬ ‫ݱ‬
Improvement Shrinkage ߝ ଴ ሺ‫ݐ‬ሻ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬
Compliance ‫׎‬ሺ‫ݐ‬ǡ ‫ ݐ‬ᇱ ሻ ‫ݱ‬ ‫ݱ‬ ‫ݱ‬
Humidity (h(t) ‫ݱ‬ ‫ݱ‬ ‫ݱ‬
History
Temperature T(t) ‫ݱ‬ ‫ݱ‬ ‫ݱ‬
Compliance ‫׎‬ሺ‫ݐ‬ǡ ‫ ݐ‬ᇱ ሻ ‫ݱ‬
Direct Shrinkage ߝ ଴ ሺ‫ݐ‬ሻ ‫ݱ‬
Strength fcyl (t) ‫ݱ‬

10.21.3.Slab dimensions

All details and geometry of the slab specimens in the ATENA-Creep analysis are similar to
the models in ATENA-Static. In addition, the applied element type for concrete, reinforcement
and loading pads in creep analysis are similar to those in static analysis. Based on the sensitivity
analysis of the time-dependent deflection to the mesh size, a minimum of 8 elements per

435
Chapter 10- Numerical analysis

thickness of the slab are selected to mesh the slab structure. Mesh size of the pads is in
agreement with the mesh size of the main concrete structure.

10.21.4.Monitoring points for long-term deflection

The point at mid-span of the slab is selected as the monitoring point in creep analysis. The
time-dependent deflection variation under the effects of creep and shrinkage over 500 days is
verified by the experimental results. Moreover, development of the cracking pattern along the
slab span is monitored under the loading levels and simultaneous effects of creep and shrinkage.

10.21.5.Loading steps

The creep analysis in ATENA contains two types of applied loads. The first type is the
sustained service load, and the second type is the creep implementation in the model. To avoid
any loading shock and the unexpected effects of sudden loading on the time-dependent
behaviour of slabs, application of the loads on the model was accomplished in multiple steps.
Creep or other non-linear effects of concrete will cause a redistribution of stresses in the
structure. For an accurate processing, a sufficient number of time steps is required. The time
spacing is created by the number of the time steps. Determination of the time steps and time
spacing depends on the type of the analysed structures as well as on the choice of time units.

Figure (10.34) shows the input table in the ATENA-Creep environment to define the interval
data for loading types and to assign the proper number of the loading steps for each interval.

436
Chapter 10- Numerical analysis

Figure 10.34: Input data table to define the interval data in creep analysis

The results of creep analysis can be plotted by the maximum deflection of the slab at mid-
span versus the age of concrete structure after loading. Figure (10.35) shows a typical
deflection-time diagram in the ATENA-Creep environment.

Figure 10.35: Typical deflection-time diagram of LWC slab in ATENA environment

437
Chapter 10- Numerical analysis

10.21.6.Verification of the proposed models

Based on the experimental results, the models presented in Equations (6.5), (6.7) and (6.8) of
Chapter 6 have been developed in this study to predict the shrinkage and creep of lightweight
concrete containing expanded polystyrene beads respectively. Details of the utilized data and
verification of the developed models have been explained in Chapter 6 of this study.

The proposed models for creep and shrinkage behaviour are verified by comparing the
experimental and numerical analysis of deflection-time diagrams in the FE models.

Figure (10.36) compares the experimental deflection-time diagram in loading level 1 (LL1)
with predictions of the model, determined by utilizing the proposed creep and shrinkage models.

40
Proposed model
35
Experimental LL1
30
Deflection (mm)

25

20

15

10

0
0 50 100 150 200 250 300 350 400 450 500
Time (Days)

Figure 10.36: comparison of the proposed model with the experimental data in loading level 1

Figure (10.37) compares the experimental deflection-time diagram in loading level 2 (LL2)
with predictions of the model, determined by utilizing the proposed creep and shrinkage models.

438
Chapter 10- Numerical analysis

40
Proposed model
35
Experimental LL2
30
Deflection (mm)

25

20

15

10

0
0 50 100 150 200 250 300 350 400 450 500
Time (Days)

Figure 10.37: Comparison of the proposed model with experimental data in loading level 2

The diagrams in Figures (10.36) and (10.37) confirm that the proposed creep and shrinkage
models are in a good agreement with the measured experimental data in two loading levels. The
predicted deflections at early ages in both figures are less than the recorded experimental values.
This trend of lower predicted values continues up to the age of 350 days. After 350 days, the
experimental deflection goes down to the predicted values. However, the long-term deflections
after 500 days in both loading levels are very close to the experimental values.

439
Chapter 10- Numerical analysis

10.21.7.Plane stress and plane strain analysis

Some problems in elasticity may be treated reasonably by a two-dimensional or plane theory


of elasticity. Plane stress and plane strain are two general types of problems involved in the
plane analysis and are defined by setting down specific restrictions and assumptions on the
stress and displacement fields and their physical prototypes.

As shown in Figure (10.38a), the plane stress is defined to be a state of stress in which the
normal stress, σz, and the shear stresses, τxz and τyz, directed perpendicularly to the x-y plane are
assumed to be zero. The geometry of the body is essentially that of a plate with one dimension
much smaller than the others. The loads are applied uniformly over the thickness of the plate
and act in the plane of the plate. The plane stress condition is the simplest form of behaviour of
continuum structures and represents situations frequently encountered in practice.

As shown in Figure (10.38b), the plane strain is defined to be a state of strain in which the
strain normal to the x-y plane, εz, and the shear strain γxz and γyz are assumed to be zero. In the
plane strain, one deals with a situation in which the dimension of the structure in one direction
is enormous in comparison to the dimensions of the structure in the other two directions. The
geometry of the body is fundamentally that of a prismatic cylinder with one dimension much
larger than the others. The applied forces act in the x-y plane and do not vary in the z direction,
i.e. the loads are uniformly distributed with respect to the large dimension and act
perpendicularly to it. Some important practical applications of plane strain condition occur in
the analysis of dams, tunnels, and other geometrical works. Also, many small-scale problems
such as bars and rollers compressed by forces normal to their cross section are amendable to
analysis under plane strain condition.

a b
Figure 10.38: Plane stress (a) and plane strain (b) presentation of a general element in the structure

Table (10.15) illustrates the general assumptions in plane stress and plane strain conditions of
the structure.

440
Chapter 10- Numerical analysis

Table 10.15: Stress and strain assumptions in plane stress and plane strain conditions

Plane stress Plane strain


Stress ߪ௭ ൌ Ͳ߬௫௭ ൌ Ͳ߬௬௭ ൌ Ͳ ߬௫௭ ൌ Ͳ߬௬௭ ൌ Ͳ
“σ” ߪ௫ ǡ ߪ௬ ܽ݊݀߬௫௬ ƒ›Šƒ˜‡‘œ‡”‘˜ƒŽ—‡• ߪ௫ ǡ ߪ௬ ǡ ߪ௭ ܽ݊݀߬௫௬ ƒ›Šƒ˜‡‘œ‡”‘˜ƒŽ—‡•
Strain ߛ௫௭ ൌ Ͳߛ௬௭ ൌ Ͳ ߝ௭ ൌ Ͳߛ௫௭ ൌ Ͳߛ௬௭ ൌ Ͳ
“ε” ߝ௫ ǡ ߝ௬ ǡ ߝ௭ ǡ ߛ௫௬ ƒ›Šƒ˜‡‘œ‡”‘˜ƒŽ—‡• ߝ௫ ǡ ߝ௬ ǡ ߛ௫௬ ƒ›Šƒ˜‡‘œ‡”‘˜ƒŽ—‡•

This study investigates the long-term deflection analysis of LWC slabs in 2D and 3D
environments in the ATENA program. Considering the symmetry of the slab geometry, support
conditions, reinforcement arrangement and loading, it is reasonable to analyse the slab in the 2D
environment to get acceptable results. However, the proper selection of plane stress or plane
strain elements may affect the accuracy of the analysis.

Generally, 2D analysis takes one meter thickness of the structure in third dimension by
default, while the width of the investigated slabs is 400 mm in the third dimension. This study
investigates the effect of plane stress or plane strain elements in the third dimension to find out
the most effective depth of structure in 2D creep analysis of the structure.

Comparing the plane strain and plane stress elements in analysing the slab specimens, and
considering the geometry of the slabs and direction of the applied load on top surface of the
slabs, plane stress elements give a better simulation of the flexural deflection of slabs under two
different loading levels.

10.21.7.1. Effect of depth of structure in 2D analysis


To study the sensitivity of the time-dependent deflection to the thickness of plane stress
element, different values of thickness have been implemented in the input data, and the model
has been analysed.

The creep and shrinkage values were maintained unchanged during the analyses. Also the
compliance function of the LWC mixture in consideration of the relaxation of concrete is
included in the models.

Figure (10.39) shows a typical input data table in the GID-ATENA program for creep
analysis. As mentioned earlier, the default value for element thickness in the third dimension (z
direction) is 1.0 metre. The thickness of elements in x and y directions are defined by mesh size
in 2-D analysis.

441
Chapter 10- Numerical analysis

Figure 10.39: Typical input data table for creep analysis in GID-ATENA

- Loading level 1

This section contains the results of analysis of the models by different thickness values of
plane stress or plane strain. The models are subjected to the first level of the uniformly
distributed sustained loading equal to 30% of the ultimate moment capacity.

Although the plane strain element is not compatible with the slabs behaviour under the
flexural loading in the 2-D environment, to have the comparison base with the model with plane
stress element, some analyses have been accomplished by plane strain element.

Figure (10.40) compares the time-dependent deflection versus time in the model with plane
strain elements of various thickness values. The experimental data is well-predicted by the
models with 0.6 to 0.8 mm thickness.

442
Chapter 10- Numerical analysis

80
0.1 m 0.3 m
70 0.4 m 0.6 m
0.8 m 1.0 m
60 Experimental L1
Deflection (mm)

50

40

30

20

10

0
0 50 100 150 200 250 300 350 400 450 500
Time (Days)
Figure 10.40: Deflection versus time in the model with different thickness values of plane strain element

Figure (10.41) shows time-dependent deflection of the slabs versus time in 2D analysis by
applying the plane stress elements with different thickness values. The applied load level is
equivalent to 30% of the ultimate bending capacity of the slabs.

60
0.2 m 0.3 m
0.4 m 0.5 m
0.6 m 0.7 m
50 0.8 m 1.0 m
Experimental L1

40
Deflection (mm)

30

20

10

0
0 50 100 150 200 250 300 350 400 450 500
Time (Days)
Figure 10.41: Deflection versus time in the model with different thickness values of plane stress element

The experimental result of long-term deflection of slab subjected to first loading level lies
between the predicted curves by thicknesses of 0.4 to 0.8m, while the exact depth of structure in
the z direction is 0.4 m. In other words, the thickness of 2-D models strongly affects the
443
Chapter 10- Numerical analysis

predicted trend of deflection and the maximum value of the time-dependent deflection after 500
days. Also, the 2-D model gives higher values of the time-dependent deflection for exact
thickness (0.4 m) of the structure in the z direction.

Figure (10.42) compares the ratio of the maximum deflection after 500 days in the plane
stress models to the experimental values. The higher thickness of the plane stress element gives
fewer values of the deflection after 500 days in the model.

2.5
Deflection ratio at 500 days

1.5

0.5

0
0.2

0.3

0.4

0.5

0.6

0.7

0.8

1
Thickness of plane stress element (m)

Figure 10.42: Ratio of deflection of plane stress element models to the experimental deflection at 500 days

- Loading level 2

This section contains the results of analysis of the models by different thickness values of
plane stress or plane strain subjected to the second level of the uniformly distributed sustained
loading equal to 40% of the ultimate moment capacity of slab section.

Figure (10.43) shows the time-dependent deflection of the slabs in 2-D analysis by applying
the plane stress element with different thickness values.

The experimental result of long-term deflection of the slab subjected to second loading level
lies between the predicted curves with thickness values of 0.3 to 0.75 m. In other words, the
thickness of 2D models strongly affects the predicted trend and maximum value of the time-
dependent deflection after 500 days. Also, the 2-D model gives higher values of the time-
dependent deflection for exact thickness (0.4m) of the structure in z direction.

444
Chapter 10- Numerical analysis

80
0.15 m 0.2 m
70 0.3 0.4
0.6 m 0.75 m
1.0 m 0.5 m
60
Experimental
Deflection (mm) 50

40

30

20

10

0
0 50 100 150 200 250 300 350 400 450 500
Time (Days)
Figure 10.43: Deflection versus time in the model with different thickness values of plane stress element

Figure (10.44) compares the ratio of the maximum deflection after 500 days in the plane
stress models to the experimental values. The higher thickness of the plane stress element gives
fewer values of the deflection after 500 days in the model.

2.5
Deflection ratio at 500 days

1.5

0.5

0
0.2

0.3

0.4

0.5
0.6

0.9
1
0.45
0.15

0.25

0.35

0.75

Thickness of plane stress lement (m)

Figure 10.44: Ratio of deflection of plane stress element models to the experimental deflection at 500 days

According to Figures (10.43) and (10.44), increasing the thickness of the plane stress element
in 2D analysis, gives lower values of the maximum deflection of LWC slab at mid-span. In
other words, the lesser depth of structure in the third dimension in 2D analysis enlarges the
resultant time-dependent deflection of slabs for 500 days.

445
Chapter 10- Numerical analysis

Figure (10.45) compares the ratio of the maximum deflection at 500 days for different
thickness values of the plane stress elements under two loading levels. The figure shows the
importance of the loading level in creep analysis of the plane stress elements.

The ratio of the corresponding deflection in each thickness varies in the range of 1.12 to 1.27.
On average, the deflection at loading level 2 is about 16% higher than the corresponding value
in loading level 1.

60
Max. deflection after 500 days (mm)

L1
50 L2
40

30

20

10

0
0.2 0.4 0.6 0.8 1
Thickness of plane stress element (m)

Figure 10.45: Effect of loading level and thickness of plane stress element on maximum deflection

10.21.8.Comparing plane stress, plane strain and 3D elements

There are different types of elements used to model LWC slabs in ATENA. Figure (10.46)
compares the time-dependent trend of deflection over 500 days by implementing plane stress,
plane strain and 3D elements. The thickness of elements in plane stress and plane strain
elements in 2D analysis is taken as 0.4 m. Initial values of the deflection of the model with
plane stress and plane strain elements are very close to the recorded experimental values,
however, the trend and the maximum deflection after 500 days are higher. On the other hand,
despite lower initial values of deflection by implementing 3D elements, the time-dependent
deflection after 500 days is very close to the experimental value. However, the increasing rate of
deflection under combined effects of creep and shrinkage is entirely different from the
experimental results.

446
Chapter 10- Numerical analysis

45
Plane stress Plane strain
40
3D Experimental L1
35
30
Deflection (mm) 25

20

15
10

0
0 50 100 150 200 250 300 350 400 450 500
Time (Days)
Figure 10.46: Comparing the plane stress, plane strain and 3D elements

10.22. EFFECT OF THE IMPLEMENTED CREEP MODELS ON THE


LONG-TERM DEFLECTION OF SLABS

This section compares the implemented creep models in the ATENA program to predict the
time-dependent deflection of a LWC slab under the creep and shrinkage effects. The existing
models in the ATENA program are the commonly used models in the engineering applications
of concrete and reinforced concrete structures. However, they have been developed originally
for the conventional concrete other than LWC containing expanded polystyrene beads. The
environmental conditions, physical and mechanical properties of the concrete and the mixture
proportioning are other crucial parameters in developing those creep models.

This study aims to discover the most compatible creep model with the experimental results of
LWC slabs containing expanded polystyrene beads. The selected model(s) can be modified for
different conditions, properties and aggregate types to develop a more specific model for LWC.
Also, the compatible models can be applied to verify the proposed creep and shrinkage models
and the models for mechanical properties of LWC, presented in Chapters 5 and 6. Utilized
models in the ATENA-creep problem type are presented in Table (10.3)

The ATENA-Creep problem type enables the user to include or exclude the effects of the
compliance (see section 4.2 in Chapter 4) and shrinkage of the concrete in or from the creep
analysis. It makes feasible studying of the individual and combined effects of the shrinkage and
compliance on the time-dependent behaviour of the concrete structure feasible.

Figures (10.47 a, b) show the options of activating the compliance and shrinkage effects in
the implemented creep models in the ATENA programs. Both the compliance and shrinkage

447
Chapter 10- Numerical analysis

activation option (Figure 10.47 a) is included in the FIBMC2010, B3, B3 improved and the BP-
KX creep models of ATENA. However, in CEB-FIP78, CSN-731201, BP1 and BP2 creep
models, only the shrinkage effect activation is included (Figure 10.47 b), and the compliance
effect is included in the creep data. A comprehensive list of the implemented creep models in
ATENA and the effective parameters in each model are previously presented in in section
(10.21) and Table (10.14) of this Chapter.

(a) (b)
Figure 10.47: Options to activate the shrinkage and compliance effects in creep analysis in ATENA

10.22.1.B3 creep model of ATENA

Figures (10.48) and (10.49) compare the deflection-time diagrams by implementing the B3
creep model of ATENA and the experimental data in the slabs subjected to the loading levels 1
(LL1) and loading level 2 (LL2) respectively. Furthermore, the individual effects of the
shrinkage and compliance of concrete are analysed in the models.

By comparing the diagrams in Figure (10.48), the shrinkage has a stronger effect than the
compliance on the time-dependent deflection of LWC slab. Also, the simultaneous effect of the
creep and shrinkage on B3 creep model gives a lower prediction of the time-dependent
deflection in the slab subjected to LL1. The porous structure of the lightweight concrete gives
higher values of the shrinkage strain, while the B3 creep model in Table (10.14) is not
completely compatible with the lightweight concrete in terms of including the air content, slump
and aggregate ratio. In other words, the shrinkage data are the actual experienced data, while the
implemented creep data in the model is not compatible with the utilized.

Regarding the shrinkage and compliance effects, diagrams in Figure (10.49) for LL2 are in
good agreement with the diagrams of LL1 shown in Figure (10.48); i.e. the effect of shrinkage in

448
Chapter 10- Numerical analysis

the time-dependent deflection of slab is higher than the effect of compliance of the concrete in
both LL1, LL2.

In B3 creep model without shrinkage and compliance effects in Figure (10.49), the ultimate
deflection at 500 days is approximately equal to the instantaneous deflection under the effect of
the applied loading on the slab.

30
Shrinkage, No compliance
Compliance, No shrinkage
25 Shrinkage and compliance
Experimental L1

20
Deflection (mm)

15

10

0
0 50 100 150 200 250 300 350 400 450 500
Time (Days)
Figure 10.48: B3 creep model with and without shrinkage and compliance effects in loading level 1

40
No shrinkage no compliance
Shrinkage, No compliance
35 Compliance, No shrinkage
Shrinkage and compliance
30 Experimental L2
Deflection (mm)

25

20

15

10

0
0 50 100 150 200 250 300 350 400 450 500
Time (Days)
Figure 10.49: B3 creep model with and without shrinkage and compliance effects in loading level 2

449
Chapter 10- Numerical analysis

10.22.2.FIBMC-2010 creep model of ATENA

Figures (10.50) and (10.51) compare the prediction of deflection-time diagrams by the
FIBMC-2010 creep model of ATENA and the experimental data of the slabs subjected to LL1
and LL2 respectively. Moreover, the individual effects of the shrinkage and compliance of
concrete are compared in the models. The less compatibility of the creep model with the utilized
concrete type, and the application of real shrinkage data is the main difference in the resulted
diagram.

From the comparison of the diagrams shown in Figure (10.50), the shrinkage poses a greater
effect than the compliance on the time-dependent deflection of LWC slab. Also, the
simultaneous effect of the creep and shrinkage in the FIBMC-2010 model gives greater
prediction of the time-dependent deflection in the slab under LL1.

Concerning the shrinkage and compliance effects, diagrams in Figure (10.51) for LL2 are in
reasonable agreement with the diagrams in Figure (10.50) for LL1; i.e. the effect of shrinkage on
the time-dependent deflection of the slab is higher than the effect of compliance of the concrete
in both LL1 and LL2.

Also, the simultaneous effect of creep and shrinkage in the FIBMC-2010 model gives higher
prediction of the time-dependent deflection for the slabs under LL1 and LL2.

The FIBMC-2010 creep model without shrinkage and compliance effects in Figures (10.50)
and (10.51) gives greater deflection than the instantaneous deflection of slabs under both LL1
and LL2.

40
Shrinkage, No compliance Compliance, No shrinkage
35 Shrinkage and compliance Experimental L1
No shrinkage, no compliance
30
Deflection (mm)

25

20

15

10

0
0 50 100 150 200 250 300 350 400 450 500
Time (Days)
Figure 10.50: FIBMC-2010 creep model with and without shrinkage and compliance effects in loading level 1

450
Chapter 10- Numerical analysis

45
No shrinkage no compliance
Shrinkage, No compliance
40 Compliance, No shrinkage
Shrinkage and compliance
35 Experimental L2

30
Deflection (mm) 25

20
15

10

0
0 50 100 150 200 250 300 350 400 450 500
Time (Days)
Figure 10.51: FIBMC-2010 creep model with and without shrinkage and compliance effects in loading level 2

10.22.3.BP-KX creep model of ATENA

Figures (10.52) and (10.53) compare the prediction of deflection-time diagrams by the BP-
KX creep model of ATENA and the experimental data of slabs subjected to LL1 and LL2
respectively. In addition, the individual effects of the shrinkage and compliance are compared in
the model.

Regarding the shrinkage and compliance effects, the diagrams shown in Figure (10.53) for
LL2 are in good agreement with the diagrams shown in Figure (10.52) for LL1; i.e. the effect of
shrinkage on the time-dependent deflection of the slab is higher than the effect of compliance of
the concrete in both LL1 and LL2. Also, the simultaneous effect of the creep and shrinkage in the
this model gives a higher prediction of the time-dependent deflection for the slabs under LL1
and LL2.

The BP-KX creep model of ATENA without shrinkage and compliance effects in Figures
(10.52) and (10.53) gives higher deflection compared to the instantaneous deflection of slabs
under both LL1 and LL2. Furthermore, effect of the compliance of concrete on the deflection-
time diagram is not constant with time.

451
Chapter 10- Numerical analysis

40
Shrinkage and compliance Shrinkage, No compliance
35 Compliance, No shrinkage Experimental L1

30 NO shrinkage, No compliance

25
Deflection (mm)
20

15

10

0
0 50 100 150 200 250 300 350 400 450 500
Time (Days)
Figure 10.52: BP-KX creep model with and without shrinkage and compliance effects in loading level 1

45
No shrinkage no compliance Shrinkage, No compliance
40 Compliance, No shrinkage Shrinkage and compliance
Experimental L2
35

30
Deflection (mm)

25

20
15

10

0
0 50 100 150 200 250 300 350 400 450 500
Time (Days)
Figure 10.53: BP-KX creep model with and without shrinkage and compliance effects in loading level 2

10.22.4.CEB-FIP78 creep model of ATENA

Figures (10.54) and (10.55) compare the predicted deflection-time diagrams by the CEB-
FIP78 creep model with the experimental data of the slabs subjected to LL1 and LL2
respectively. In addition, the individual effect of the shrinkage is evaluated in the model.

From the comparison of the diagrams shown in Figure (10.54), inclusion of the shrinkage
effect in the analyses changes the deflection-time diagrams for LL1 and LL2 considerably.
However, the effect of shrinkage in LL1 and LL2 is different.

The deflection-time diagram of the slab by the CEB-FIP78 creep model, including the
shrinkage effect, is in moderate agreement with the experimental results, especially for LL1.

452
Chapter 10- Numerical analysis

Despite a similar trend of time-deflection, the increasing rate of deflection after 250 days is
considerably higher than the experimental data for LL2.

30
Shrinkage

25 No shrinkage

Experimental L1
20
Deflection (mm)

15

10

0
0 50 100 150 200 250 300 350 400 450 500
Time (Days)

Figure 10.54: CEB-FIP78 creep model with and without shrinkage effect in loading level 1

50
Shrinkage
45
No shrinkage
40
Experimental L2
35
Deflection (mm)

30
25
20
15
10
5
0
0 50 100 150 200 250 300 350 400 450 500
Time (Days)

Figure 10.55: CEB-FIP78 creep model with and without shrinkage effect in loading level 2

10.22.5.CSN-731201 creep model of ATENA

Figures (10.56) and (10.57) compare the predicted deflection-time diagrams by the CSN-
731201 creep model of ATENA with the experimental data of the slabs subjected to LL1, LL2
respectively. Also, the individual effect of the shrinkage is studied in the model.

453
Chapter 10- Numerical analysis

The comparison of the diagrams shown in Figure (10.56) and (10.57), including the shrinkage
effect in the FEA, significantly changes the deflection-time diagrams in LL1and LL2
respectively. However, the effect of shrinkage in LL1 and LL2 is different.

The deflection-time diagram shown in the slab by CSN-731201 creep model inclusion of the
shrinkage effect is approximately in agreement with the experimental results in LL2. However,
the predicted deflection in LL1 is noticeably below the experimental data.

In the first loading level shown in Figure (10.56), despite a similar trend of time-deflection
diagrams up to 350 days, the model with the shrinkage effect changes the deflection behaviour.
However, the ultimate deflation is substantially below the exact experimental value.

30
Shrinkage

25 No shrinkage

Experimental L1
20
Deflection (mm)

15

10

0
0 50 100 150 200 250 300 350 400 450 500
Time (Days)

Figure 10.56: CSN-731201 creep model with and without shrinkage effect in loading level 1

40
Shrinkage
35
No shrinkage
30 Experimental L2
Deflection (mm)

25

20

15

10

0
0 50 100 150 200 250 300 350 400 450 500
Time (Days)

Figure 10.57: CSN-731201 creep model with and without shrinkage effect in loading level 2
454
Chapter 10- Numerical analysis

10.22.6.BP1 creep model of ATENA

Figures (10.58) and (10.59) compare the predicted deflection-time diagrams by the BP1 creep
model of ATENA with the experimental data of slabs subjected to LL1, LL2 respectively. In
addition, the individual effect of the shrinkage is considered in the model.

From the comparison of the diagrams shown in Figure (10.58) and (10.59), inclusion of the
shrinkage effect in the analysis changes the deflection-age diagrams, especially in LL2
significantly. However, the effect of shrinkage in LL1 and LL2 is totally different.

The deflection-time diagram of the slab by the BP1 creep model, including the shrinkage
effect, is in high agreement with the experimental data in LL1. This creep model gives
reasonable predictions in LL2.

35
Shrinkage
30 No shrinkage

Experimental L1
25
Deflection (mm)

20

15

10

0
0 50 100 150 200 250 300 350 400 450 500
Time (Days)
Figure 10.58: BP1creep model with and without shrinkage effect in loading level 1

40
Shrinkage
35
No shrinkage
30 Experimental L2
Deflection (mm)

25

20

15

10

0
0 50 100 150 200 250 300 350 400 450 500
Time (Days)
Figure 10.59: BP1creep model with and without shrinkage effect in loading level 2
455
Chapter 10- Numerical analysis

10.22.7.BP2 creep model of ATENA

Figures (10.60) and (10.61) compare the predicted deflection-time diagrams of the BP2 creep
model of ATENA with the experimental data of slabs subjected to LL1 and LL2 respectively.
Additionally, the individual effect of the shrinkage is analysed in the model.

From the comparison of the diagrams shown in Figures (10.60) and (10.61), inclusion of the
shrinkage effect in the FEA, strong changes in the deflection-time diagrams in both LL1 and LL2
are clear. However, the effect of the shrinkage in LL1 and LL2 is different.

The deflection-time diagram in the slab of the BP2 creep model including the shrinkage
effect is approximately in agreement with the experimental results in LL2. However, the
predicted deflection in LL1 is considerably below the experimental data.

In the first loading level shown in Figure (10.60), despite a similar trend of the time-
deflection diagrams up to 420 days, the model with shrinkage effect changes the deflection
behaviour. However, the ultimate deflations are significantly below the exact experimental
values.

30
Shrinkage

25 No shrinkage

Experimental L1
20
Deflection (mm)

15

10

0
0 50 100 150 200 250 300 350 400 450 500
Time (Days)

Figure 10.60: BP2 creep model with and without shrinkage effect in loading level 1

456
Chapter 10- Numerical analysis

40
Shrinkage
35
No shrinkage
30 Experimental L2

Deflection (mm)
25

20

15

10

0
0 50 100 150 200 250 300 350 400 450 500
Time (Days)

Figure 10.61: BP2 creep model with and without shrinkage effect in loading level 2

10.22.8.Comparing the creep models of ATENA

Figures (10.62) and (10.63) compare different creep models used to predict the deflection-
time diagrams under the two loading levels of LL1 and LL2 respectively. Unpredictably, the
loading level influences the predictions of these creep models. In other words, the models are
highly sensitive to the applied sustained loading. However, in the ordinary engineering projects,
to determine the creep coefficient, only specified stress level is applied on the test specimens.

As shown in Figure (10.62), the FE models of the slabs on the basis of the FIBMC-2010,
CEB-FIP78 and BP1 creep models, give the best estimation of time-dependent deflection of the
slabs subjected to LL1. However, the predicted values and diagrams by the BP2 and CSN-
731201 creep model significantly underestimate the experimental deflection of the LWC slabs
under LL1.

457
Chapter 10- Numerical analysis

35
B3 FIBMC-2010

30 BP1 BP2
CSN-731201 CEBFIP-78
25 BPKX Experimental L1

Deflection (mm) 20

15

10

0
0 50 100 150 200 250 300 350 400 450 500
Time (Days)

Figure 10.62: Comparing the implemented creep models in loading level 1

The FEA with all the creep models except CEBFIP-78, predict the long-term deflection
values close to the experimental values under LL2 in Figure (10.63). However, the CSN-731201
creep model reveals the best agreement with the experimental data. Furthermore, the BP1 and
BP2 models in the simulated slab models are in good agreement with the experimental data.

50
B3 FIBMC-2010
45 BP1 BP2
40 CSN-731201 CEBFIP-78
BPKX Experimental L2
35
Deflection (mm)

30
25
20
15
10
5
0
0 50 100 150 200 250 300 350 400 450 500
Time (Days)

Figure 10.63: Comparing the implemented creep models in loading level 2

Comparing the results in Figures (10.62), (10.63) confirms the sensitivity of the creep models
to the sustained loading level in the prediction of the time-dependent deflection of slabs. Among
the implemented creep models in ATENA, BP2 is strongly influenced by the load level, while
FIBMC-2010 shows the least variation from the experimental results under two loading levels.
458
Chapter 10- Numerical analysis

Variation of the deflection by other creep models is between BP2 and FIBMC-2010. They are
between the extreme points of the compatibility with the experimental data.

As general conclusion, FIBMC-2010 is the most compatible model for the time-dependent
behaviour of LWC slab, therefore it is recommended to apply these particular creep and
shrinkage effects in study of the structures with this type of concrete.

10.23. SUMMARY

The proposed models for the mechanical and time-dependent properties of lightweight
concrete and the deflection-age behaviour of slabs need to be verified by finite element analysis.
Among the existing finite element programs, the ATENA software was selected to model and
analyses of the slabs. ATENA has the capability to implement the non-linear and time-
dependent behaviours that are crucial in this study. A brief description of the existing
capabilities of ATENA in relation to the accurate modelling of the reinforced concrete
structures, together with the material models for concrete, reinforcement and content elements
were given. The element types and compatibility between the material and element types were
also given to assign the most appropriate element and meshing to the selected materials.

Two types of problems in ATENA, namely ATENA-Static and AATENA-Creep were


utilized to study the instantaneous and long-term deflection of the lightweight concrete slabs
respectively. In each problem type, a comprehensive parametric study was performed to
calibrate the simulated models to show the most compatible results with the experimental data.
The following results can be drawn from the finite element analysis of the calibrated models in
ATENA in the static and creep problem types.

- The finite element models verify the proposed models of mechanical properties of
lightweight concrete in Chapter Five;
- The finite element models verify the proposed creep and shrinkage models of lightweight
concrete in Chapter Six;
- 3D analysis is the most accurate method, however, regarding the symmetric geometry of
slabs and loading and supporting conditions, 2D analysis can be used with high accuracy
of the results both in static and creep analysis;
- In 2D analysis, the plane stress analysis of the models shows better agreement with the
experimental data;
- The smaller mesh size and higher number of the loading steps increase the accuracy of
the e results both in static and creep analysis;

459
Chapter 10- Numerical analysis

- The loading level has a considerable effect on the obtained results, particularly on the
creep analysis. In other words, loading level can change the creep effect on the structure;
- The slab model were analysed by applying different creep models and compared with the
experimental data. The loading level changed the effect of each creep model;
- Time-dependent analysis of the models shows the higher effect of the shrinkage strain
than compliance of the concrete, on the long-term deflection of concrete slabs;
- In the first loading level (0.3Mu), FIBMC-2010, CEB-FIP78 and BP1 creep models give
the best estimation of the time-dependent deflection of lightweight concrete slabs;
- In the second loading level (0.4Mu), CSN-731201 creep model gives the best estimation
of the time-dependent deflection of lightweight concrete slabs. Overall, FIBMC-2010 is
the most compatible model for the time-dependent behaviour of LWC slab;
- Among the creep models, BP2 is strongly influenced by the load level, while FIBMC-
2010 shows the least variation with the experimental results under two loading levels.

Figures (10.64) and (10.65) show the compatibility between the experimental data and
predictions of the implement creep models.

40
Proposed model
35
Experimental LL1
30
Deflection (mm)

25

20

15

10

0
0 50 100 150 200 250 300 350 400 450 500
Time (Days)

Figure 10.64: comparison of the proposed model with the experimental data in loading level 1

460
Chapter 10- Numerical analysis

40
Proposed model
35
Experimental LL2
30
Deflection (mm)

25

20

15

10

0
0 50 100 150 200 250 300 350 400 450 500
Time (Days)

Figure 10.65: Comparison of the proposed model with experimental data in loading level 2

461
Chapter 10- Numerical analysis

REFERENCES

1. Ngo, D. and A. Scordelis. Finite element analysis of reinforced concrete beams. in ACI Journal
Proceedings. 1967. ACI.
2. Chaudhari, S.V. and M.A. Chakrabarti, Modeling of concrete for nonlinear analysis Using Finite
Element Code ABAQUS. International Journal of Computer Applications, 2012. 44(7): p. 14-18.
3. Yuanfeng, L.X.W., Three-dimensional nonlinear finite element analysis of reinforced concrete
structures based on ANSYS program. 2nd International Conference on Computer Engineering
and Technology, 2010. 6: p. 42-46.
4. Smadi, M. and K. Belakhdar, Development of finite element code for analysis of reinforced
concrete slabs. Jordan Journal of Civil Engineering, 2007. 1(2).
5. Červenka, V. and L. Jendele, ATENA Program Documentation Part 2-2 User’s Manual for
ATENA 3D. 2014.
6. ABAQUS, ABAQUS MANUAL 2010.
7. Lee, H.H., Estimation of Uncertainties and Validation of Computational Models for Structural
Concrete. 2009: ProQuest.
8. Hu, H.-T. and W.C. Schnobrich, Constitutive modeling of concrete by using nonassociated
plasticity. Journal of Materials in Civil Engineering, 1989. 1(4): p. 199-216.
9. Gilbert, R.I. and Z.I. Sakka, Strength and ductility of reinforced concrete slabs containing
welded wire fabric and subjected to support settlement. Engineering Structures, 2010. 32(6): p.
1509-1521.
10. Rousseau, J., P. Marin, and L. Daudeville, Discrete element modeling of reinforced concrete
with a particular steel-concrete interface. Discrete Element Group for Hazard Mitigation, 2007.
11. Chong, K.T., numerical modelling of time-dependent cracking and deformation of reinforced
concrete structures. PhD dissertation, University of New Southe Wales, Sydney, Australia,
2004.
12. Jing, H. and Y.Q. Li. Nonlinear Finite Element Analysis of Layered Steel Fiber Reinforced
Concrete Beam. in Applied Mechanics and Materials. 2012. Trans Tech Publ.
13. CERVENKA, V., J. CERVENKA, and R. PUKL, ATENA – A tool for engineering analysis of
fracture inconcrete. S¯adhan¯a, 2002. 27(4): p. 485-492.
14. Johnson, S., Comparison of Nonlinear Finite Element Modeling Tools for Structural Concrete.
2006.
15. Yang, K.-H., H.-S. Chung, and A.F. Ashour, Influence of shear reinforcement on reinforced
concrete continuous deep beams. ACI structural journal, 2007. 104(4).
16. Yu, Q., Size E®ect and Design Safety in Concrete Structures under Shear. PhD dissertation,
NORTHWESTERN UNIVERSITY, ILLINOIS, USA, 2007.
17. Malm, R. and J. Holmgren, Cracking in deep beams owing to shear loading. Part 2: Non-linear
analysis. Magazine of Concrete Research, 2008. 60(5): p. 381-388.
18. Talley, K.G., Assessment and strengthening of ASR and DEF affected concrete bridge columns.
2009.
19. Gribniak, V., et al., Finite element mesh size effect on deformation predictions of reinforced
concrete bridge girder. The Baltic Journal of Road and Bridge Engineering, 2010. 5(1): p. 19-
27.
20. Dobromil, P., C. Jan, and P. Radomir, Material model for finite element modelling of fatigue
crack growth in concrete. Procedia engineering, 2010. 2(1): p. 203-212.
21. Olteanu, I., A. Alistar, and M. Budescu, Non-linear Analysis of Reinforced Concrete Frames
with ATENA 3-D Program. Buletinul Institutului Politehnic din lasi. Sectia Constructii,
Arhitectura, 2011. 57(2): p. 93.

462
Chapter 10- Numerical analysis

22. Lampropoulos, A.P., O.T. Tsioulou, and S.E. Dritsos, Biaxial stress due to shrinkage in concrete
jackets of strengthened columns. ACI Materials Journal, 2012. 109(3).
23. Sucharda, O. and J. Brozovsky, Numerical modelling of reinforced concrete beams with
fracture-plastic material. Frattura ed Integrità Strutturale, 2014. 30: p. 375-382.
24. Singh, V., et al., Finite Element Modeling of CFRP Retrofitted RC Beam-Column Joints.
International Journal on Emerging Technologies, 2014. 5(2): p. 31.
25. Meskenas, A., et al., Simplified technique for constitutive analysis of SFRC. Journal of Civil
Engineering and Management, 2014. 20(3): p. 446-453.
26. Namy, M., J.-P. Charron, and B. Massicotte, Structural behavior of cast-in-place and precast
concrete barriers subjected to transverse static loading and anchored to bridge deck overhangs.
Canadian Journal of Civil Engineering, 2015. 42(2): p. 120-129.
27. Kollegger, J. and G. Mehlhorn, Experimentelle und Analytische Untersuchungen zur Aufstellung
eines Materialmodels fuer Gerissene Stahbetonscheiben. Nr. 6 Forschungsbericht, Massivbau,
Gesamthochschule Kassel, 1988.
28. Dyngeland, T., Behaviour of Reinforced Concrete Panels. Doktor Ingenierafhandling (Ph. D.
thesis, Institut for Betongkonstruksjoner, Trondheim, 1989.
29. béton, C.e.-i.d., CEB-FIP model code 1990: design code. 1993: CEB Bulletin d’Information
No.213/214, Comité EuroInternational du Béton, Lausanne, Switzerland, Telford.
30. Bigaj, A.J., Structural dependence of rotation capacity of plastic hinges in RC beams and slabs.
1999.
31. Jensen, E. and W. Hansen, Nonlinear aggregate interlock model for concrete pavements.
International Journal of Pavement Engineering, 2006. 7(4): p. 261-273.
32. Yang, K.-H. and A.F. Ashour, Aggregate interlock in lightweight concrete continuous deep
beams. Engineering Structures, 2011. 33(1): p. 136-145.
33. Choi, C.-K. and H.-G. Kwak, The effect of finite element mesh size in nonlinear analysis of
reinforced concrete structures. Computers & Structures, 1990. 36(5): p. 807-815.
34. Zheng, Y., Finite element analysis for fixture stiffness. 2005.

463
Chapter 11: Summary and conclusions

CHAPTER 11

SUMMARY AND CONCLUSIONS


Chapter 11: Summary and conclusions

Table of Contents

11.1. SUMMARY............................................................................................................................ 464


11.2. CONCLUSIONS .................................................................................................................... 466
11.2.1. Experimental program; Hardened Concrete Properties .......................................................... 467
11.2.2. Time-dependent properties of hardened concrete ................................................................... 469
11.2.3. Experimental program; Flexural Load-Deflection Behaviour ................................................ 470
11.2.4. Experimental program; Flexural Instantaneous deflection ..................................................... 472
11.2.5. Experimental program; flexural long-term deflection ............................................................ 473
11.2.6. Numerical investigations ........................................................................................................ 477
11.3. RECOMMENDATIONS FOR FUTURE RESEARCH ......................................................... 480
11.3.1. Mechanical properties: ....................................................................................................... 480
11.3.2. Structural behaviour ........................................................................................................... 480
11.3.3. Especial applications .......................................................................................................... 481
Chapter 11: Summary and conclusions

List of Figures

Figure 11.1: Comparison of deflection vs. time in SCC, LWC and CC slabs under lower loading level 475
Figure 11.2: Comparison of deflection vs. time in SCC, LWC and CC slabs under higher loading level475
Figure 11.3: Predicted Δins and Δtot of SCC, LWC and CC slabs under two different loading levels ...... 476
Figure 11.4: comparison of the proposed model with the experimental data in loading level 1 .............. 478
Figure 10.5: Comparison of the proposed model with experimental data in loading level 2 ................... 479
Chapter 11: Summary and conclusions

List of Tables

Table 11.1: Creep coefficient of CC, LWC and SCC mixtures after 240 days sustained loading ........... 470
Table 11.2: Summary of the results from short-term flexural test ........................................................... 471
Table 11.3: Loading values, recorded and elastic deflection of SCC, CC and EPA-LWC slabs ............. 472
Table 11.4: Ratio of the long-term to instantaneous deflection of LWC, SCC and CC slabs ................. 474
Chapter 11: Summary and conclusions

Chapter 11: Summary and conclusion

11.1. SUMMARY

This study investigated the deflection behaviour of lightweight reinforced concrete simply-
supported one-way slabs containing treated expanded polystyrene beads. In addition, the
deflection behaviour of slabs was monitored in three distinct categories as a) incremental
deflection under increasing load until the flexural failure of slab, b) instantaneous deflection at
the time of loading and c) time-dependent deflection under sustained loading.

Regarding the time-dependent deflection, a particular emphasis was on the development of


the flexural deflection attributed to the combined effects of the constant service loads and the
time-dependent properties of concrete such as creep and shrinkage. The investigation included
both the experimental and analytical parts to verify the results precisely. For the analytical part,
a finite element analysis using the ATENA program and statistical methods were utilized.

The overall objective of this study was to compare the deflection behaviour of the slabs made
of different types of concrete. Nejadi (2005) conducted an experimental study on the cracking
and deflection behaviour of conventional concrete slabs for 400 days. Aslani (2014)
investigated the time-dependent cracking and deflection behaviour of self-compacting concrete
slabs for 250 days.

Details of the specimen geometry (length, thickness, and depth), support conditions, and
reinforcement details in the experiments are the same for all three series of conventional, self-
compacting and lightweight concrete slabs. The applied loading on the lightweight and self-
compacting concrete slabs were 30% and 40% of the ultimate bending capacity of the slabs

464
Chapter 11: Summary and conclusions

section. However, for conventional concrete slabs, the uniformly distributed loading was about
20% and 30% of the ultimate bending capacity of the slabs section.

The deflection history and development with age were the main parameters in monitoring the
deflection behaviour. The key drivers of the deflection development were also investigated. The
fresh properties of concrete, bond characteristics, the time-dependent effects of creep and
shrinkage on the lightweight concrete members thereof have been also monitored and studied.

The properties of concrete in the hardened state, including the mechanical properties,
shrinkage and creep and the bond-slip between the reinforcing bars and the surrounded concrete
were also studied. The hardened properties of lightweight concrete have been compared to the
corresponding properties of the conventional concrete and self-compacting concrete.
Furthermore, new models have been proposed as a means of predicting the mechanical
properties of concrete in comparison with the existing models in the literature.

To study the effect of short-term loading on the deflection behaviour of slabs, a series of tests
was performed on the simply-supported one-way lightweight concrete slabs. The slabs were
subjected to monotonically increasing loads up the failure point to construct the load-deflection
diagram and determine the ultimate loading capacity. The deflection of each slab at mid-span,
crack patterns, steel tensile strains, and concrete surface strains at the level of the reinforcing
bars were carefully monitored under the increasing load up to the failure point.

Mechanical properties of the light-weight concrete including the compressive strength, the
modulus of rupture, the modulus of elasticity, the bond strength between the concrete and steel
bars and the splitting tensile strength of lightweight concrete were measured at different ages.
Also, the creep and shrinkage strain of the lightweight concrete were measured by the
companion specimens.

Another series of tests were conducted to evaluate the long-term deflection of lightweight
concrete slabs. In this regard, four simply supported one-way slabs were subjected to two
distinct levels of sustained loads for 525 days. Development of the flexural deflection under the
combined effects of service load and creep and shrinkage effects were also studied. Deflection
of slabs at mid-span and the point on the border of the high-moment region, crack patterns, steel
tensile strains and the concrete surface strains at the level of reinforcing bars were recorded
immediately after loading for up to 525 days under two constant values of the sustained load on
each pair of slabs. Moreover, the compression strain of concrete at the top surface on mid-span
of the slab was monitored under the loadings.

The finite element program was utilized to model the slab structures to simulate the
instantaneous and time-dependent flexural cracking and deflection. New analytical models were

465
Chapter 11: Summary and conclusions

proposed to predict the mechanical properties of lightweight concrete. These models were
verified by the obtained results from the numerical models of the slabs in the ATENA program.

The results of the experimental tests were compared with predictions of the proposed models
and the existing models in the literature. The comparison included both instantaneous and long-
term deflections.

A comprehensive parametric study for both types of the experimental study (instantaneous
and time-dependent deflection) of the slabs by finite element analysis has been performed. The
parametric analysis of the models in finite element analysis included the proposed models for
hardened properties, bond characteristics, creep, and shrinkage models.

11.2. CONCLUSIONS

This study is presented in eleven chapters. Chapters One to Five contain the literature review
of the used material, structural concepts of the topic and relevant investigations. In Chapters
One the research methodology and layout of the thesis are described. Chapter Two includes a
brief history of the invention, development and mechanical properties of lightweight concrete.
In Chapter Three, experimental results of almost all investigations about the lightweight
concrete containing expanded polystyrene beads and its hardened properties are collected and
analysed. Chapter Four describes the time-dependent properties of concrete and existing models
of creep, shrinkage, and steel-concrete bond characteristic. In chapter Five, mechanism of the
flexural deflection and effective parameters and existing models are broadly discussed. In
addition, in the literature review, main differences and existing models of the mechanical
properties of self-compacting concrete, conventional concrete and lightweight concrete are
comprehensively evaluated.

Experimental and numerical phase of the study are explained in Chapters Six to Ten. Chapter
Five includes the hardened properties of lightweight concrete. Chapter Seven explain the
flexural load-deflection of slabs. In Chapters Eight and Nine, the instantaneous and Long-term
flexural deflections of slabs are presented. Chapter Ten explains the verification of experimental
investigation by finite element analysis. Summary and conclusions of the experimental and
numerical phase of the study are presented in Chapter Eleven in the following main categories:

- Hardened concrete properties;


- Flexural load-deflection of one-way slabs;
- Flexural Instantaneous deflection of one-way slabs;
- Flexural long-term deflection of one-way slabs;
- Numerical analysis of one-way slabs.
466
Chapter 11: Summary and conclusions

11.2.1. Experimental program; Hardened Concrete Properties

11.2.1.1. Material properties

New models are proposed in this study to estimate the mechanical properties of lightweight
concrete including the Compressive Strength (CS), Modulus of Elasticity (MoE), Modulus of
Rupture (MoR), Splitting Tensile Strength (STS), Compressive Stress-Strain Curve (CSSC) and
energy absorption capacity. In addition, CS, MoE, MoR and STS of Self-Compacting Concrete
(SCC), Conventional Concrete (CC) and Light-Weight Concrete (LWC) and their development
with time are compared. Form the proposed models and comparison of the results, the following
conclusions can be made:

- Generally, the achievable compressive strength is considered as the main difference


between SCC, CC and LWC. However, despite different nature of mix design and
proportions of SCC, CC and LWC, it is possible to reach a relatively high compressive
strength in these types of concrete.
- The existing models of MoE, MoR, STS and CSSC for LWC in the literature are based
on the specific conditions and particular type of lightweight aggregates in different
researches. The proposed models for MoE, MoR, STS and CSSC for LWC containing
Expanded Poly-Styrene (EPS) beads in this study attempt to solve this problem.
- Despite similar range of the compressive strength of the investigated SCC, CC and
EPS-LWC, there is considerable difference between the other mechanical properties.
- The early-age properties and development rate in SCC, especially, the SCC mixture
reinforced with different types of fibre are higher than CC and LWC. Development rate of
CS, MoR and MoE in EPS-LWC with time is the least among all concrete types.
- The presence of PES beads in the used LWC, makes the CSSC different form the SCC
and CC. The relatively brittle failure of EPS-LWC gives sharper descending branch of CSSR
compared to that of SCC and CC. However, the ascending branch of CSSC in EPS-LWC is
similar to that of the investigated CC and SCC.
- Due to brittle failure EPS-LWC compared to CC and SCC, the area under the stress-
strain curve is smaller; therefore, the absorbed energy is less than that of SCC and SCC in
the same grade compressive strength.
- The proposed models for MoE, MoR, STS and CSSR in this study that are on the base
of the most compatible models presented in Chapter Three, are in good agreement with the
experimental data. Some modifications have been applied to the best-matching existing
models to get the best results of the proposed models.

467
Chapter 11: Summary and conclusions

11.2.1.2. Bond-slip between the embedded bar and surrounding concrete

The following conclusions can be made from the pull-out test in this study and comparison of
the results with other types of concrete and literature.
- The bond failure surface through the lightweight aggregates particularly in low-strength
LWC, while in CC specimens most aggregates are not broken. For higher strength
levels over 40 MPa, the cracking surface is similar in both types of concrete;
- Considering SCC, CC and LWC, the ultimate bond strength and mean bond strength are
the highest and lowest values in SCC and LWC respectively.
- Corresponding bond strength in LWC is relatively lower than that for CC. However, in
high-strength LWC, the bond strength is greater than that of CC, due to higher mortar
strength.
- Embedded length, type and diameter of steel bar in concrete, compressive strength of
concrete, concrete cover, shape and size of the concrete specimen and aggregate type
are the most effective parameters in the bond-slip behaviour of SCC, CC and LWC.
However, due to lower values of compressive strength, effects of other parameters are
more obvious in LWC.
- The maximum bond stress and bond-slip behaviour are presented in many empirical
models and codes of practice. However, they are not accurate enough to predict the
maximum bond stress and bond-slip relationship in EPS-LWC.
- The most compatible existing models with the experimental data of EPS-LWC are
modified to propose new models to get the best compatibility with the experimental
data. The proposed models of the maximum bond stress and bond-slip accurately is in
agreement with the experimental data.
- Due to lower bond stress between the embedded bar in EPS-LWC cylindrical
specimen in this study, there is considerable slip seen between the ribbed embedded
bars and surrounding EPS-LWC. In most cases, the brittle failure of EPS-LWC
specimen due to tension force ended the pull-out test. However, in some cases, there
were no cracking in the cylinder even after pulling out the steel bar.
- The ultimate failure force in pull-out test is increasing by age of specimen. However,
due to the improvement of the bond-strength by time, there is less slip of steel bar seen
by increasing the age of the specimen.

468
Chapter 11: Summary and conclusions

11.2.2. Time-dependent properties of hardened concrete

11.2.2.1. Shrinkage of EPS-LWC


According to the test results and discussions presented in Chapter Six regarding the shrinkage
behaviour of EPS-LWC, and comparison with the shrinkage strain of SCC and CC, the
following conclusions can be made:
- The ultimate shrinkage strain in EPS-LWC is considerably higher than that of SCC and
CC. The shrinkage strain in SCC and CC are similarly in the range of 800-900 με, while
it reaches to about 1200 με.
- The Shrinkage strain of EPS-LWC is strongly influenced by the shape and size of the
specimen. The cylinder specimen showed about 15% higher ultimate shrinkage strain
than prism specimens for the same mixture and similar curing and laboratory
conditions.
- Based on the results of different curing, drying and rewetting conditions of the
shrinkage specimens of EPS-LWC, the duration of initial curing is the most effective
parameter in shrinkage of EPS-LWC. The longer the initial curing time, the smaller the
ultimate shrinkage strain.
- Rewetting of specimens followed by a period of drying after initial curing time shows
almost no improving effect on the shrinkage of EPS-LWC.
- The existing models in ACI-435-R95, AS-3600-09 and ACI-209-08 provide
underestimated shrinkage strain of EPS-LWC. The predicted ultimate shrinkage strain
in ACI-209-08, AS-3600-09 and ACI-435-R95 is about 33%, 50% and 50% of the
measured experimental shrinkage strain of EPS-LWC respectively.
- The empirical shrinkage model proposed by Best and Polivka (1959) gives the best
prediction of the shrinkage strain of EPS-LWC compared to the other existing models.
- The proposed model in this study, accurately predicts the shrinkage strain of EPS-LWC.
- There is a remarkable relationship between the rate of weight loss (water loss) of
shrinkage specimen and time. The model proposed in this study to accurately predict the
weight loss of EPS-LWC specimen at different ages.

11.2.2.2. Creep of EPS-LWC

According to the test results and comparison of the creep strain and creep coefficient in EPS-
LWC, SCC and CC, the following conclusions can be made:

- The creep behaviour of all types of concrete is highly sensitive to the level of sustained
stress. In average, 10% higher sustained stress caused about 25% higher creep strain in
the concrete specimen;

469
Chapter 11: Summary and conclusions

- Since, in most design codes and standards, the design creep is determined from a
specified magnitude of sustained stress, application of the determined creep for other
stress levels should be verified;
- The creep cylindrical specimens in SCC, CC and EPS-LWC were loaded at age 14
days. The creep coefficient in EPS-LWC is considerably higher than that of SCC and
CC cylindrical specimens. While, the CC showed the least creep coefficient. Table
(11.2 shows the creep coefficient of SCC, CC and EPS-LWC specimens;
-
Table 11.1: Creep coefficient of CC, LWC and SCC mixtures after 240 days sustained loading

CC LWC N-SCC D-SCC S-SCC DS-SCC

Creep coefficient 1.5 2.415 1.96 1.86 1.82 1.45

- Despite the similar trend of creep increment at first days in all concrete types, the
growing rate of creep strain in EPS-LWC after about 50 days is noticeably higher than
the creep increment of SCC and CC specimens;
- The existing creep models provide underestimated creep behaviour of EPS-LWC.
However, the most compatible models have been selected and modified to propose new
accurate models;
- The models proposed in this study, precisely predict the creep strain and creep
coefficient of EPS-LWC. In the proposed models to predict the creep strain and creep
coefficient of EPS-LWC, the effect of the sustained stress level is also included.

11.2.3. Experimental program; Flexural Load-Deflection Behaviour

The short-term load increment until the failure and corresponding deflection values at mid-
span of slabs were monitored. From the experimental results and comparison of the load-
deflection behaviour in SCC, CC and EPS-LWC slabs, the following conclusions can be made:

Table (11.2) shows the failure load and the maximum deflection of the slab at mid-span in the
short-term failure loading.

- The slab geometry (span length and section dimensions) and reinforcement details in all
SCC, CC and EPS-LWC slabs were similar. In addition, despite the similar range of the
compressive strength of the slabs, there was a significant difference in the modulus of
elasticity of applied SCC, CC and EPS-LWC mixtures in the slabs.
- Considering the ratio of tensile reinforcement in the section, all slabs experienced
ductile failure under the increasing flexural loading. In other words, the ultimate
470
Chapter 11: Summary and conclusions

bending capacity of the slab section was more sensitive to the yield strength of the steel
bars rather than the compressive strength of the concrete.
- As presented in Table (11.2), there is no considerable difference in the failure loads of
SCC, CC and EPS-LWC slabs and all failure loads are in the range of 50 KN. While
- Despite the similar range of the failure load in all slabs, there is a substantial difference
in the maximum deflection of slabs at mid-span at failure load.
- Fibre reinforcing of the SCC mixture improves the ductility of the slabs. The ultimate
flexural deflection of these slabs is considerably higher than the other slabs.
- Compared to SCC and CC slabs, the lower value of the modulus of elasticity in EPS-
LWC slabs resulted in a smaller amount of the maximum deflection at failure load. The
relatively brittle failure of the EPS-LWC slabs may be considered as the reason of the
difference in the recorded ultimate deflection.

Table 11.2: Summary of the results from short-term flexural test

Failure Deflection at
Failure Deflection at
Slabs load Failure load
Beams load (KN) failure load (mm)
(KN) (mm)

NSCC-A 49 180 LWC-1 46.5 95.9


concrete slabs
Self-compacting concrete slabs

Lightweight

NSCC-B 48.5 163 LWC-2 48 88

D-SCC-A 53 205 LWC-3 41.53 114.7

D-SCC-B 52 177 LWC-4 49.52 86.65


S-SCC-A 50 185
concrete slabs
Conventional

S-SCC-B 48 167 S3-a 50 136


DS-SCC-A 56 220 S3-b 47 156
DS-SCC-B 54 182

- There is a noticeable difference in cracking pattern of the slabs under increasing load in
terms of the crack spacing and crack width. The average crack spacing in SCC and CC
slabs are less than in EPS-LWC slabs. However, the crack height and crack width in
SCC and CC slabs are higher than those in EPS-LWC slabs.
- As a general conclusion of the load-deflection of slabs, the failure load, and
corresponding maximum deflection should be considered simultaneously in the design
of one-way slabs.

471
Chapter 11: Summary and conclusions

11.2.4. Experimental program; Flexural Instantaneous deflection

The flexural deflection of slabs immediately after loading at age 14 days was recorded. In this
regard, three different types of reinforced concrete slabs made with normal, lightweight and
self-compacting concrete slabs was investigated under two distinct levels of service loads.From
the experimental results and comparison of the flexural instantaneous deflection of SCC, CC
and EPS-LWC slabs, the following conclusions can be made:

- The concrete type strongly affects the deflection behavior of the reinforced concrete
slabs;
- Elastic deflection is not a trustworthy indicator of the instantaneous, time-dependent and
total deflection of slabs;
- Considering the elastic deflection (Δe) equal to instantaneous deflection (Δins) even for
service load under the cracking load in slabs is completely wrong and causes unsafe
design of reinforced concrete slabs to calculate the time-dependent and long-term
deflection.
- The ratio of measured instantaneous deflection to calculated elastic deflection at 14 days
varies from 1.14 to 3.8 in one-way slabs.

Table (11.3) shows the applied uniformly distributed service load (wa), service moment (Ma),
calculated elastic deflection at 14 and 28 days and the measured instantaneous deflection of
slabs at 14 days.

Table 11.3: Loading values, recorded and elastic deflection of SCC, CC and EPA-LWC slabs

wa Ma Δ e (mm) Δ ins
Slab
(KN/m) (KN.m) 14 day 28 day (mm)
N-SCC-a 7.31 11.189 3.18 2.9 12.1
N-SCC-b 5.41 8.28 2.36 2.15 5.89
D-SCC-a 7.26 11.12 3.5 2.85 7.65
D-SCC-b 5.36 8.21 2.58 2.11 7.59
S-SCC-a 7.3 11.18 3.45 2.87 6.41
S-SCC-b 5.4 8.27 2.55 2.12 2.91
DS-SCC-a 7.33 11.23 3.29 2.85 8.98
DS-SCC-b 5.43 8.32 2.44 2.11 5.14
LWC-1 7 10.72 4.99 4.57 11.96
LWC-2 7 10.72 4.99 4.57 11.4
LWC-3 5.29 8.095 3.77 3.45 6.3
LWC-4 5.29 8.095 3.77 3.45 5.98
CC-a 4.44 6.8 2.73 2.5 11.8
CC-b 3.46 5.3 2.13 1.95 5.04

- Effective moment of inertia (Ie) is the most critical parameter in calculating the flexural
instantaneous deflection. In majority of the existing models to predict the effective

472
Chapter 11: Summary and conclusions

moment of inertia, effects of applied service moment (Ma), cracking moment (Mcr), the
gross moment of inertia (Ig) and cracking moment of inertia (Icr) are included. This study
confirms the implementation of the loading age factor and ultimate moment capacity of
the reinforced concrete section. In fibre reinforced sections, the effect of fibre volume in
the mixture should be considered in the Ie model.
- The ratio of service moment to the ultimate moment capacity of slabs strongly affects the
instantaneous deflection;
- In the existing models in the literature, there is an almost linear relation between the load
value and the instantaneous deflection of slabs. The results of this study showed a
considerable effect of the load level, especially in the low strength concrete slabs;
- The instantaneous deflection due to 10% of load increment increased about 100%, and
105% in the lightweight concrete and self-compacting concrete slabs. In low-strength
normal concrete slabs, only 6% higher load resulted in 135% higher instantaneous
deflection;
- The instantaneous deflection of fiber reinforced self-compacting concrete slabs
experienced fewer variations with the changes of loading level;
- Development of the mechanical properties of normal, lightweight and self-compacting
concrete is different before the age of 28 days. Therefore, for the early age loading, the
calculations based on the mechanical properties at the age of 28 days may result in the
wrong values for the instantaneous deflection of the slabs;
- Almost all existing Ie models in the literature estimate higher values of the instantaneous
deflection for the investigated slabs. The most compatible models in the literature are
selected as the basis of new proposed models in this study. Implementation of the loading
age, ratio of Modification on the age

- A new model is proposed and verified to predict the effective moment of inertia in the
self-compacting concrete slabs with and without fiber reinforcing. The predicted
instantaneous deflections by this model are in agreement with the experimental results;
- Another model of the effective moment of inertia is proposed and verified for the normal
and lightweight concrete slabs. The experimentally recorded instantaneous deflections are
well-predicted by this model.

11.2.5. Experimental program; flexural long-term deflection

The flexural deflection of slabs under sustained service loads and effects of creep and
shrinkage was recorded. In this regard, three different types of reinforced concrete slabs made

473
Chapter 11: Summary and conclusions

with SCC, CC and EPS-LWC slabs were investigated under two distinct levels of service loads.
From the experimental results and comparison of the flexural long-term deflection of SCC, CC
and EPS-LWC slabs, the following conclusions can be made:

- The concrete type strongly affects the time-dependent deflection behavior of reinforced
concrete slabs;
- In EPS-LWC slabs, the time-dependent deflection was monitored at mid-span and the
point on the border of the high moment region. The general trend of the time-dependent
deflection of at these points was similar. However, the ratio of deflection at mid-span to
the deflection at the border point of high moment region is considerably different in the
slabs under different loading levels.
- The long-term deflection is considered as the multiplier of the instantaneous deflection in
the majority of design codes. Table (11.4) shows the measured deflections of SCC, CC
and EPS-LWC slabs at the time zero after loading (instantaneous deflection) and the final
deflection after long-term sustained loading.
- Regardless of the loading level, the deflection multiplier in design codes varies in the
range of 2 to 3. According to Table (11.4), this ratio is in agreement with the long-term
deflection of slabs at lower loading level. By 10% increase in the loading level, the
deflection multiplier increases about 50% above the range of design codes.
-

Table 11.4: Ratio of the long-term to instantaneous deflection of LWC, SCC and CC slabs

Ma / Mu (%) Test duration Δ tot Δ ins


Slab Δ tot / Δ ins
14 day 28 day (days) (mm) (mm)
N-SCC-a 40.73 40.27 240 24.76 12.1 2.04
N-SCC-b 30.14 29.8 240 18.08 5.89 3.07
D-SCC-a 39.94 39.65 240 17.76 7.65 2.32
D-SCC-b 29.48 29.27 240 16.78 7.59 2.21
S-SCC-a 40.32 39.85 240 22.27 6.41 3.47
S-SCC-b 29.83 29.48 240 20.25 2.91 6.95
DS-SCC-a 40.03 39.63 240 21.31 8.98 2.37
DS-SCC-b 29.66 29.36 240 15.83 5.14 3.08
LWC-1 39.13 38.78 525 34.56 11.96 2.89
LWC-2 39.13 38.78 525 31.81 11.4 2.79
LWC-3 29.55 29.28 525 26.4 6.3 4.19
LWC-4 29.55 29.28 525 24.52 5.98 4.1
CC-a 26.12 25.14 400 32.1 11.8 2.72
CC-b 20.36 19.59 400 21.92 5.04 4.35

474
Chapter 11: Summary and conclusions

Figures (11.1) and (11.2) compare the mid-span deflection versus time in the SCC, CC and
EPS-LWC slabs under the lower and higher loading levels, respectively. The loading level of
each slab is shown in Table (11.4).

30

25
Deflection (mm)

20

15

10 LWC-D D-SCC-B DS-SCC-B

N-SCC-B S-SCC-B CC-B


5

0
0 50 100 150 200 250
Age (days)
Figure 11.1: Comparison of deflection vs. time in SCC, LWC and CC slabs under lower loading level

30

25
Deflection (mm)

20

15

10
LWC-A LWC-B D-SCC-A

5 DS-SCC-A N-SCC-A S-SCC-A


CC-A

0
0 50 100 150 200 250
Age (days)
Figure 11.2: Comparison of deflection vs. time in SCC, LWC and CC slabs under higher loading level
475
Chapter 11: Summary and conclusions

Figures (11.3 a, b) compare the predictions of AS-3600 (2009), ACI-318 (2008) and
CSA.A23 (2004) for the instantaneous and total (long-term) deflection of SCC, LWC and CC
slabs under two different levels of sustained loading respectively.

- The instantaneous and total deflection in the slabs under higher loading level is
significantly greater than those of the slabs under lower loading levels, especially by
ACI-318 and CSA.A23 predictions. However, in predicted values of AS-3600, the
difference between the Δins-14 and Δlong is less, compared to the other design codes.
- According to Table (11.4), the experimental ratio of Δtot / Δins-14 varies between 2.04 and
4.35; however, the investigated codes of practice predict the ratio in the range of 1.89 to
2.31.

50
45 SCC-a LWC-a CC-a
Predicted deflection (mm)

40
35
30
25
20
15
(a)
10
5
0
Δins Δtot Δins Δtot Δins Δtot
ACI-318 AS-3600 CSA.A23
Codes of practice

35
SCC-b LWC-b CC-b
30
Predicted deflection (mm)

25

20

15
(b)
10

0
Δins Δtot Δins Δtot Δins Δtot
ACI-318
AS-3600 CSA.A23
Codes of practice
Figure 11.3: Predicted Δins and Δtot of SCC, LWC and CC slabs under two different loading levels

476
Chapter 11: Summary and conclusions

- There is a considerable difference between the predicted Δins-14 and Δlong of slabs under
different load values. The instantaneous and total deflection in the slabs under higher
loading level is significantly greater than those of the slabs under lower loading levels,
especially by ACI-318 and CSA.A23 predictions. However, in predicted values of AS-
3600, the difference between the Δins-14 and Δtot is less, compared to the other design
codes.
- All listed codes in Figures (11.3 a, b) predict the highest values of Δins-14 and Δtot for EPS-
LWC slabs under both loading levels. However, AS-3600 predicts the lowest Δins-14 and
Δtot values for SCC slabs under both loading levels.
- By comparing the ratio of Δtot / Δins-14 in Table (11.4) and the drying shrinkage strain of
SCC-CC and EPS-LWC, there is no obvious relationship between the ratio of Δtot / Δins-14
and the shrinkage behaviour of slabs.
- The rate of deflection (mm/day) and the free shrinkage rate (με/day) in SCC, CC and
EPS-LWC are not compatible.

11.2.6. Numerical investigations

Two types of problems in ATENA, namely ATENA-Static and ATENA-Creep were utilized
to study the instantaneous and long-term deflection of the EPS-LWC slabs respectively. In each
problem type, a comprehensive parametric study was performed to calibrate the simulated
models to show the most compatible results with the experimental data. The following results
can be drawn from the finite element analysis of the calibrated models in ATENA in the static
and creep problem type:

- The proposed models of mechanical properties of EPS-LWC in Chapter Six are verified;
- The finite element models verify the proposed creep and shrinkage models of lightweight
concrete in Chapter Six;
- 3D analysis is the most accurate method, however, regarding the symmetric geometry of
slabs and loading and supporting conditions, 2D analysis can be used with high accuracy
of the results both in static and creep analysis;
- In 2D analysis, the plane stress analysis of the models shows better agreement with the
experimental data;
- The smaller mesh size and higher number of the loading steps increase the accuracy of
the results both in static and creep analysis;
- The loading level has a considerable effect on the obtained results, particularly on the
creep analysis. In other words, loading level can change the creep effect on the structure;

477
Chapter 11: Summary and conclusions

- The slab models were analysed by applying different creep models and compared with
the experimental data. The loading level changed the effect of each creep on the slab
deflection;
- Time-dependent analysis of the models shows the higher effect of the shrinkage strain
than compliance of the concrete, on the long-term deflection of concrete slabs;
- In the first loading level (0.3Mu), FIBMC-2010, CEB-FIP78 and BP1 creep models give
the best estimation of the time-dependent deflection of lightweight concrete slabs;
- In the second loading level (0.4Mu), CSN-731201 creep model gives the best estimation
of the time-dependent deflection of lightweight concrete slabs. Overall, FIBMC-2010 is
the most compatible model for the time-dependent behaviour of LWC slab;
- Among the creep models, BP2 is strongly influenced by the load level, while FIBMC-
2010 shows the least variation with the experimental results under two loading levels.

Figures (11.4) and (11.5) show the compatibility between the experimental data and
predictions of the proposed creep and shrinkage models in this study.

40
Proposed model
35
Experimental LL1
30
Deflection (mm)

25

20

15

10

0
0 50 100 150 200 250 300 350 400 450 500
Time (Days)

Figure 11.4: comparison of the proposed model with the experimental data in loading level 1

478
Chapter 11: Summary and conclusions

40
Proposed model
35
Experimental LL2
30

Deflection (mm)
25

20

15

10

0
0 50 100 150 200 250 300 350 400 450 500
Time (Days)

Figure 10.5: Comparison of the proposed model with experimental data in loading level 2

479
Chapter 11: Summary and conclusions

11.3. RECOMMENDATIONS FOR FUTURE RESEARCH

The following areas of research regarding the lightweight concrete members remain relatively
unexplored and could form the basis for future studies:

11.3.1. Mechanical properties:


- The structural lightweight concrete members are excellent solutions for lighter structures
with higher performance. However, fibre reinforcing can increase the performance and
mechanical properties of lightweight structures while the gravity loads and the resultant
lateral effects are almost unchanged. A wider investigation is required to develop the
flexural failure mechanism of fibre reinforced lightweight concrete to use in beams and
slabs.
- There are many sources of natural and artificial lightweight aggregates to make the
lightweight concrete. Comparing the mechanical properties of lightweight concrete with
different types of lightweight aggregates to achieve a similar structural concrete (density
and compressive strength) might be an interesting topic in choosing the most economic
and accessible aggregates to make the lightweight concrete;
- Considering the higher thermal and sound insulation, and better energy absorption of
lightweight concrete, an experimental program to compare the lightweight and the new
types of construction materials might be interesting in the construction industry.

11.3.2. Structural behaviour


- The current study emphasized the flexural capacity of the lightweight concrete structures,
while other failure mechanisms such as shear, axial and torsional failures are not taken
into account. Considering the shear failure importance in most engineering structures,
more investigations are needed to develop an accurate failure mechanism to describe the
collapse of lightweight concrete slabs with and without fibre reinforcement.
- Surprisingly, experimental studies about the time-dependent properties of lightweight
concrete are very rare in the literature. The long-term experimental program can help to
discover the uncertain parameters in the serviceability of lightweight concrete members
with and without fibre reinforcement.
- Load-bearing capacity, fatigue and the effects of cyclic and repeating loads in short or
long-term durations in the structural lightweight concrete still require more studies.

480
Chapter 11: Summary and conclusions

11.3.3. Especial applications


- An extensive experimental and numerical program is required to investigate the long-
term behaviour of precast prestressed lightweight concrete members;
- Lightweight concrete containing expanded polystyrene beads is frequently used in marine
structures and decks;
- Application of nanotechnology to modify the mixture design and structural performance
of the lightweight concrete containing expanded polystyrene beads is almost new and
requires more investigation;
- Lightweight concrete is a type of green materials in regard to the lower weight and
consequently reduced gas emission, reduced pollution and fewer mining and logistics
problems. The comparison of the energy saving and environmental considerations in
building with conventional and different types of lightweight concrete requires further
studies;
- Sustainability and durability are the main factors in the application of the concrete in
most of the structures. Studying the freeze-thaw resistance, chloride penetration and other
chemical attacks particularly in the absence of harsh conditions experienced by marine
structures requires empirical and numerical investigation.

481
Appendix A: Material tests

APPENDIX A

MATERIAL TESTS

482
Appendix A: Material tests

PREPARATION OF TEST SPECIMENS

483
Appendix A: Material tests

484
Appendix A: Material tests

485
Appendix A: Material tests

486
Appendix A: Material tests

PULL OUT TEST

487
Appendix A: Material tests

488
Appendix A: Material tests

489
Appendix A: Material tests

490
Appendix A: Material tests

SPLITTING TENSILE STRENGTH

491
Appendix A: Material tests

492
Appendix A: Material tests

MODULUS OF RUPTURE

493
Appendix A: Material tests

494
Appendix B- Manufacturing the slabs and test setup

APPENDIX B

MANUFACTURING THE SLABS AND TEST


SETUP FOR LOAD-DEFLECTION TEST,
INSTANTANEOUS DEFLECTION
MONITORING AND LONG-TERM
DEFLECTION MONTORING

495
Appendix B- Manufacturing the slabs and test setup

496
Appendix B- Manufacturing the slabs and test setup

497
Appendix B- Manufacturing the slabs and test setup

498
Appendix B- Manufacturing the slabs and test setup

499
Appendix B- Manufacturing the slabs and test setup

500
Appendix B- Manufacturing the slabs and test setup

501
Appendix B- Manufacturing the slabs and test setup

502
Appendix B- Manufacturing the slabs and test setup

503
Appendix B- Manufacturing the slabs and test setup

504

You might also like